Vous êtes sur la page 1sur 611

Theory of flow and fracture of solids

Nádai, Arpád, 1883-


New York, McGraw-Hill, 1950-[63]

http://hdl.handle.net/2027/uc1.b4140760

Public Domain, Google-digitized


http://www.hathitrust.org/access_use#pd-google

We have determined this work to be in the public domain,


meaning that it is not subject to copyright. Users are
free to copy, use, and redistribute the work in part or
in whole. It is possible that current copyright holders,
heirs or the estate of the authors of individual portions
of the work, such as illustrations or photographs, assert
copyrights over these portions. Depending on the nature
of subsequent use that is made, additional rights may
need to be obtained independently of anything we can
address. The digital images and OCR of this work were
produced by Google, Inc. (indicated by a watermark
on each page in the PageTurner). Google requests that
the images and OCR not be re-hosted, redistributed
or used commercially. The images are provided for
educational, scholarly, non-commercial purposes.
B M mD 7L.Q
ft
'
ENGINEERING SOCIETIES MONOGRAPHS
Ralph H. Phelps, Consulting Editor

THEORY OF FLOW
AND

FRACTURE OF SOLIDS
ENGINEERING SOCIETIES
MONOGRAPHS
national engineering societies, the American
Four
Society of Civil Engineers, American Institute of
Mining and Metallurgical Engineers, The American

Society of Mechanical Engineers, and American Institute


of Electrical Engineers, have made arrangements with the

McGraw-Hill Book Company, Inc., for the production of

selected books adjudged to possess usefulness for engineers

or industry, but not likely to be published commercially

because of too limited sale without special introduction.

The societies assume no responsibility for any statements

made in these books. Each book before publication has,

however, been examined by one or more representatives of

the societies competent to express an opinion on the merits

of the manuscript.

ENGINEERING SOCIETIES MONOGRAPHS


COMMITTEE

A. S. C. E. A. I. M. E.
H. Alden Foster Reed VV. Hyde
Ernest P. Goodrich E. M. Wise

A. S. M. E. I. E. E.
A.
J. M. JURAN F. Malcolm Farmer
T. R. Olive W. I. Slichter

Ralph H. Phelps,
CHAIRMAN

Engineering Societies Library,


New York
To koi

THEORY OF FLOW
AND c 3

FRACTURE OF SOLIDS^"'

BY

A. NADAI
Consulting Mechanical Engineer
Weslinghouse Research Laboratories
East Pittsburgh, Pa.

Volume One
Second Edition

McGRAW-HILL BOOK COMPANY, Inc.


NEW YORK TORONTO AND LONDON

1950

WALTER T. STEILBERQ
CONSULTING ARCHITECT
NO. 1. ORCHARD LANE
BERKELEY (4) CALIF.
THORNWALL 1760
THEORY OF FLOW AND FRACTURE OF SOLIDS
(Revision of PLASTICITY)
Copyright, 1931, 1950, by the United Engineering Trustees, Inc. Printed in the
United States of America. All rights reserved. This book, or parts thereof,
may not be reproduced in any form without permission of the publishers.

THE MAPI.E PRESS COMPANY, YORK, PA.


To the memory of my dear

Elisabeth
PREFACE TO THE SECOND EDITION
Although nearly twenty years have elapsed since the first appearance
of this book, it may be said that the then newly awakened interest in a
mathematical treatment of the plastic deformations of solid materials
has not subsided. Engineers, physicists, and metallurgists desire to
have at their command the valuable experimental evidences uncovered
and assembled during the last decades in regard to the mechanical laws
of the plastic deformation and fracture of solid materials. Mathematical
theories have been broadened. In many instances progress in fields of
technology was due to, or led to, a better knowledge of the strength
properties of the metals and materials of construction.
Further developments can be expected in those sciences in which the
mechanical phenomena of the solid state are the objects of investigation.
For example, the geologists and geophysicists active in the science of
tectonics hope to find the answers to a group of interesting questions
about the causes of the deep-focus earthquakes originating at great
depths, or the remarkable regularity of the angles of dip of the cleavage
faults observed over large areas in the strata of certain rocks. These
are recognized as representing questions of a nature analogous to those
studied by engineers in the behavior of solids under combined stresses.
Whereas physicists and metallurgists interpret the problems of the solid
state by the atomic mechanisms in the lattice or grain structures of
solids, it could not be the purpose of this book to delve into these ques
tions. Passing remarks in this direction will be found scattered sparsely
in a few of the following chapters, the reader being referred in this
respect to books on the molecular theory of solids.
As already remarked in the earlier preface, investigators have tried
from time to time during the last hundred years to describe the strength
of materials beyond the elastic range by the methods of mechanics.
These efforts have not ceased but on the contrary have persistently
been renewed in the belief and hope that the disciplines of the strength
of materials might someday be based on exact definitions and treatments
comparable to those on which the mathematical theory of elasticity has
been developed.
Parts of this present volume were first published in a German mono
graph in 1927, in English in 1931, and in a Russian translation of the
American edition in 1936, and condensed parts appeared in an article
vii
viii PREFACE TO THE SECOND EDITION

in Vol. 6 of the Handbuch der Physik, Berlin, 1928. The content aims
at an improved mechanical treatment of the permanent deformation of
solids. Several new chapters offer an introduction to the theories of
simple and composite substances based on the types of strains — elastic,
permanent, or a combination of both — and on the types of laws of
deformation postulated. The treatment of the states of stress in per
manently strained cylinders and disks and the mathematical theory of
the nonhomogeneous states of plane plastic strain and of their surfaces
of slip, have been thoroughly revised. Chapters 12 and 13 add analy
ses of finite homogeneous strain based on the quadratic elongation X and
of finite plane strain utilizing expressions in terms of the components
of natural strain. A synthesis of small elastic and permanent strains
was generalized in a theory of constrained flow in the cases when the
principal axes of stress rotate relatively to the material. Whereas the
previous edition contained a treatment of the plastic deformation of
solids only, the author has endeavored to broaden the scope of the book
by attempting to survey the general conditions causing fracture in
materials and to acquaint the readers with some of the details and results
of a number of recent valuable experimental investigations on the yielding
and fracture of ductile metals under combined stress.
A number of interesting applications of the theory of plasticity could
not be covered in this volume because this would have overloaded its
content. Such a discussion is contemplated for a second volume in
which will be treated the flow of metals under concentrated pressure
with applications to the forming processes of rolling and drawing, the
theory of hardness, of residual stresses, of the forging of thin shells, of
plastic buckling of thin plates, the principle of mechanical work, and
examples of the flow of very viscous materials. The great importance
of designing machine parts which must withstand very high tempera
tures makes it imperative to include a discussion of the slow creep of
metals, devoting space to their laws of deformation. These and a few
geophysical questions concerned with related phenonema in rock strata
shall be treated in the second volume.
Similarly as this book does not propose to describe the physical or
metallurgical nature of these phenomena, considerable restraint was also
needed to guide the author in another direction, namely, in the mathe
matical treatment and presentation of the theories of flow and fracture
and of their numerous applications in important engineering problems,
assuming that such a treatment in which mathematics serves primarily
as a tool and not as an object itself may best serve the interests of a
group of readers. A thorough familiarization by the younger reader
with the geometries of the states of stress and strain is a prerequisite
PREFACE TO THE SECOND EDITION ix

for dealing with the problems the strength of materials poses. Efforts
were renewed to clarify the Mohr representation of stress and infinitesimal
strain in making a broader use also of the octahedral components of
stress and infinitesimal strain with the help of which many important
facts in the theory of plasticity have found their simplest expression.
The author hopes that engineers and physicists in their investigations
will make more use of these good means which have proved to be so
helpful for visualizing important components of the tensors and for
discussing the criteria of strength and flow in solids. One chapter is
devoted to a vector treatment of the geometries of stress and of finite
homogeneous strain in an endeavor to acquaint mathematically inclined
readers, at least, with the fundamentals of linear vector functions in the
theories of deformation of continuously distributed masses based on J.
Willard Gibbs's calculus of dyadics. One wonders that its simplicity
and perfection of form and clarity in content have not found more
recognition in books on applied mechanics. Chapter 14 follows his
" Vector Analysis" in the classic text of Edwin Bid well Wilson. Whereas
the inherent heuristic value of the calculus of dyadics for the student of
mechanics needs no recommendation, it is scarcely necessary to add
that it can offer no advantages as a means for deriving concrete solutions
of partial differential equations.
The last war increased the interest in many problems of the strength
of materials by stimulating new experimental investigations in this field.
It must be acknowledged gratefully that institutions set up by the gov
ernment during the last war had an influential part in these efforts.
The author wishes to acknowledge the interest and encouragements
offered to him in frequent discussions with the staff of the structural
mechanics section of the David Taylor Model Basin of the Navy Depart
ment in Washington. He acknowledges gratefully the value of his
frequent contacts with Captain H. E. Saunders, former director; Cap
tains W. P. Roop and J. Ormondroyd, formerly in charge of this section;
the late Dr. D. F. Windenburg, chief physicist; Drs. W. Osgood and M.
Greenfield; and others in the problems of the failure of steel plates, on
the laws of strain hardening and on the effects of high velocities of
deformation.
The author recalls also gratefully the stimulating discussions years
ago with Dr. L. Prandtl in Gottingen, and the financial support of experi
mental investigations in this period by Dr. F. Schmidt-Ott, which were
carried out there with the assistance of Drs. W. Bader, W. Lode, and
G. Mesmer, and after 1929 with the support of the late Dr. S. M. Kintner,
Vice-President of the Westinghouse Electric Corporation in Pittsburgh
and his former collaborators, Drs. C. W. MacGregor and W. 0. Rich
X PREFACE TO THE SECOND EDITION

mond. At the Westinghouse Research Laboratories the author owes


the greatest thanks to Dr. L. W. Chubb, Director Emeritus of the
Laboratories in East Pittsburgh, for his continued interest over a period
of 16 years in the problems of flow, fracture, and creep of metals and
for the assistance generously offered by him for organizing much experi
mental work; to Dr. J. A. Hutcheson for enabling me to complete this
volume; and for the help received from the drafting department. I am
much indebted to P. G. McVetty for a continued exchange of informa
tion on the creep of metals and his valuable experiences offered in this
field. To Mrs. Jolan M. Fertig, librarian at the Research Laboratories,
I am most grateful for her invaluable help during many years of assisting
me in the bibliographical work. My particular thanks are due to my
collaborators, E. A. Davis and M. J. Manjoine for their most valuable
assistance in experimental and theoretical work and for the exchange
of their experiences since 1933; to Drs. J. Miklowitz and J. Aronofsky
for recent help; to J. Getsko and Miss Jean Hoffman for their help in
the excellent preparation of the figures; to D. W. Glasser for helping
excellently to prepare many difficult photographic reproductions of test
specimens; and to my secretary, Miss Eleanor M. Lycett, for all help
received during completion of the manuscript.
A. NXdai
Westinghouse Research Laboratories
East Pittsburg, Pa.
March, 1949
PREFACE TO THE FIRST EDITION
A review of the development of the theory of elasticity will show that
during the period of over two hundred years the theory has gradually
developed into an exact part of mechanics which today is the solid
foundation for the design of engineering structures. During the last few
decades much valuable information has been obtained regarding elasti-
cally imperfect materials, the mechanical properties of ductile metals,
and the conditions of rupture in solids. However, a more satisfactory
understanding of the plastic state of engineering materials and of the
conditions of rupture has been reached since the constitution and the
crystalline structure of metals and alloys, the mechanical and thermal
history of metallic bars and sheets from the beginning of the casting of
the ingot to the last annealing of the finished product, and the mechanical
properties of metallic single crystals as special objects of investigation
have been carefully studied from a broader physical standpoint.
Considerations of the plastic state of matter are today of interest to
many branches of science and of engineering. The steel and metal
worker desires to control more accurately the mechanical processes of
forming metals at forging temperatures. Because of the large quantity
of energy at the present time consumed in steel mills during the process
of rolling, a more economical use of the energy required is needed. On
the other hand, in order to choose the right materials for various parts
of his machines, the machine designer must carefully consider the
mechanical properties of these materials. He is not only interested in
a more exact knowledge of the limiting conditions of stress at which, in
his machine parts, permanent set begins to develop and danger of yield
ing or fracture is to be expected or fatigue cracks start to form; but also
in several cases he will have to consider the possible change in shape of
machine parts exposed to long duration of stressing. He may base his
calculations as far as they refer to purely elastic deformation upon the
theory of elasticity, but he lacks in the theory of strength of materials
a similarly certain basis when considering the transition from the elastic
to plastic deformations or the conditions of rupture.
The physicist and the metallurgist are interested in the laws of plastic
flow from several points of view. The experiments carried out in recent
years with large metallic single crystals indicate clearly that plasticity
xii PREFACE TO THE FIRST EDITION

is an essential and general property of solid matter in its crystalline


state of aggregation. It has also been shown that under certain idealized
conditions plastic flow of the polycrystalline metals under low tempera
tures follows rules or laws which in their simplicity and mathematical
applicability to a variety of cases are comparable to the well-established
foundations upon which the theory of elastic deformation rests. It may
suffice to mention in this connection that since the first attempts of
de St. Venant and C. Duguet, who years ago first tried to establish a
mathematical theory of plastic deformation of metals, by the efforts of
a number of more recent mathematicians and engineers a mechanics of
the plastic state of metals has been revived with success and further
developed in many new aspects.
In minerals and rocks many evidences of plastic deformations have
been found. The changes in the structure of metals produced by plastic
deformation are in many respects analogous to certain phenomena
observed in minerals and rocks. These evidences show the correspond
ence which exists between the changes in structure in severely deformed
metals and the slow processes occurring during the formation of certain
rocks as observed and described in petrography. The conditions encoun
tered in the deeper strata of rock: long duration of small differences in
principal stress, elevated temperatures, and high average pressure, are
those favorable for producing plastic deformations in solids. To these
must be attributed some of the causes of the magnificent effects produced
by mountain building and the dislocation of the continental plates which
are observed in nature on a large scale. The remarkably regular profiles
of some of the German and North American rock-salt domes, in which a
layer of a highly plastic material such as rock salt has been pressed out
by means of mountain pressure, may be mentioned as an example where
evidences of plastic deformations are disclosed.
In the description of the plastic states of stress extensive use was made
of the surfaces of slip. The flow or slip lines, which frequently appear
as a pattern with an astonishingly regular symmetry on the surface or
in sections of solid bodies stressed above the plastic limit, have proved
an extremely interesting object of investigation and a valuable means
for analyzing the stress distributions under which they were produced.
The strange laws which seem to apply to the surfaces of slip have
attracted recently the mathematicians and the engineers and have been
discussed with success by R. v. Mises, H. Hencky, L. Prandtl, and
others. In the hands of the geologists, who in their faults have observed
similar phenomena on a large scale for a long time, these surfaces might
serve to decipher the riddles in the formation of high mountain chains,
just as their smaller relatives have helped to describe more precisely
PREFACE TO THE FIRST EDITION xiii

the plastic states in permanently deformed bodies. They might possibly


in the future still serve to study by mechanical means and to reconstruct
the history of the crustal movements of the whole continents, which,
according to the ideas of Alfred Wegener and F. B. Taylor, in their
parts or entirety are thought to drift slowly over their magmatic
substratum.
A. Nadai
Westinghouse Research Laboratories
East Pittsburgh, Pa.
July, 1931
CONTENTS
Preface to the Second Edition vii

Preface to the First Edition xi

Letter Symbols xix

PART I
DEFORMATION OF SOLIDS. ANALYSIS OF
STRESS AND STRAIN
1. Introduction 3

2. The Solid and Fluid State of Matter. Elastic, Viscous, and Plastic
Substances 10

3. Elastic and Permanent Deformation , 17

Conventional stress-strain curve — Yield stress of polycrystalline solids


depends also on rate of strain — Creep — Elastic hysteresis and aftereffects.

4. Behavior of Matter under High Pressure 30

Polymorphism — Viscosity — The compressibility of metals — The compres


sibility of artificial and natural glasses — The compressibility of liquids —
The effect of pressure on the rigidity of metals — Cohesion in liquids —
Optical and other effects — Munroe effect.

5. The Ordered and Unordered States of Matter 38

6. Crystalline Structure in Metals 43

7. Mechanism of Plastic Deformation in the Grain Structure .... 49


Slip — L. Prandtl's mechanical model of a solid —Sir Geoffrey Taylor's
theory of strain hardening of single metal crystal assuming dislocations in
atom lattice — Formation of twins — Thermal agitation in lattice of atoms,
R. Becker's theory — F. Zwicky's mosaic crystal — Creep in metals — B.
Chalmer's micro- and macrocreep in metal crystal — H. Eyring's rate
processes.

8. Theory of the Tensile Test 70

Conventional stress-strain curve — The natural stress-strain curve — Stress


distribution in neck — Mechanical similarity.

9. Stress 89

Equilibrium on tetrahedron — Determination of principal stresses — Stress


ellipsoid.
Xvi CONTENTS

10. Mohr's Representation of Stress 94


Plane stress. Stress circle. Mohr's graphical representation of a state
of stress. Shearing stress t. Principal shearing stresses. Octahedral
stresses. Deviator of stress.

11. Strain 109

Homogeneous strain — Pure extension — Simple shear — Uniform dilata


tion — Longitudinal extension without change in volume — Pure shear —
Pure rotation — Linear transformation — Infinitesimal strain.

12. Finite Homogeneous Strain without Rotation 117

Elongation X, shearing strain 7 — Graphical representation of state of finite


homogeneous strain — Graphical representation of state of finite strains
through circles — Strain components expressed through instantaneous direc
tions in strained body — Natural strain and natural shear.

13. Finite Plane Strain. Expressions Developed in Terms of the Natu


ral Strains 133

Pure shear — Pure shear expressed by natural strains — Work done under
finite pure shear. The natural shearing strains 7 and 7 — Example: Finite
pure shear in isotropic elastic material — Simple shear — Example: Finite
simple shear in isotropic elastic material.

14. Vector Geometry of Stress and Strain. Linear Vector Functions.


Tensors. Vector Fields 151

Linear vector functions — Dyads. Dyadic — Nonion and normal form of


dyadic — Symmetric, antisymmetric part of dyadic — Examples of dyadics — ■
Stress tensor — Nonion form of stress tensor and of deviator — Finite strain —
Nonion form of dyadic — Elongation X in arbitrary direction —Love's com
ponents of finite strain — Nonhomogenoous finite strain — Nabla operator.
Gradient of scalar function — Nonion form of dyadic — The condition of
incompressibility of material. The affine transformation of a sphere of
such material — Correspondence of direction cosines in unstrained and
strained condition of material.

PART II
THE YIELDING OF SOLIDS, PARTICULARLY OF THE METALS UNDER
SIMPLE STATES OF STRESS
15. Limiting States of Stress in Solids. Theories of Mechanical Strength 175
Criteria of failure — Limiting surface of yielding —Limiting surface of rup
ture — Triaxial tension — Cleavage and shear fracture — Influence of rate of
loading, time at load, stored energy in system and size of specimens on
fracture stress — Tensile strength of glass — Bursting tests with steel cylin-
ers. Herringbone fracture — Velocity of propagation of a cleavage frac
ture — Fracture along grain boundaries — Stress concentration. Effect of
notches and flaws — Fracture theory of Griffith — Fracture as an instability
of equilibrium — Fracture of heavy-walled hollow cylinders of a brittle
material under internal pressure — Theories of strength — Maximum stress
theory — Maximum elastic strain theory — Theory of constant elastic energy
CONTENTS xvii

of deformation —Theory of constant elastic strain energy of distortion or


of the constant octahedral shearing stress — Maximum shearing stress
theory — Mohr's theory of strength. J. J. Guest's condition of flow in
ductile metals — A. Leon's condition for cleavage and shear fracture — Con
dition of slip in a loose, granular material — The octahedral shearing stress
is a function of the mean normal stress <rMt.

16. Conditions of Failure. Rules in the New Theories of Flow of


Solids 229
The three rules of plastic deformation — Stress-rate of strain and stress-
strain relations.

17. Tests on Yielding and Fracture under Combined Stress 238


Testa by Guest, Foppl and Karman — The influence of the intermediate
principal stress on the yielding of metals. W. Lode's tests on ductile met
als — Tension-torsion tests by Taylor and Quinney — Ludwik's remark con
cerning the ductility of metals under biaxial states of stress — Further
tests— Plastic collapse of deep-well steel casing under external pressure
and axial tension — Tests on the strain hardening and fracture of metals
under combined stresses — Tests by E. A. Davis (coppe*, steel cylinders) —
Tests by Morikawa and Le Van Griffis (steel cylinders) — Failures of spheri
cal steel tanks and of welded cargo ships — Tests by Harmer E. Davis,
Troxell and Parker (20 in. diameter welded steel cylinders).

18. Strain and Flow Figures 275


Strain figures in mild steel— 'The- production of flow lines by notches and
holes — Elastic stress distribution and the beginning of plastic flow in a plate
with a circular hole — How plastic flow starts around a hole — Hole in region
stressed in pure shear.

19. The Yield Point of Mild Steel. Plastic Fronts. Oblique Fracture
in Flat Bars 297
The peak of the stress-strain curve of mild steel. Upper, lower yield stress.
Yield point elongation — Plastic front in a tension bar. Tests with Nylon —
Instability of tensile equilibrium in compact bars at yield point — Increase
of yield stress and yield point elongation with rate of strain — The insta
bility of tensile equilibrium in flat bars. Oblique fracture — Theory by
Bijlaard.

20. Compression 328


Cylinders — Plane compression tests — Prisms — Compression tests on cyl
inders of ductile metals — Hollow cylinders of porcelain in compression —
High strength toroid form.

21. Torsion of a Round Bar. The Stress-Strain Curve in Shear .... 347
a. Direct, b. inverse problem, c. Length changes of permanently twisted
round bars, observed by Swift and L'Hermite.

22. Plastic Bending of Bars 353


Bending of bars with arbitrary law of deformation — Pure bending of a bar
with rectangular cross section — Bar subjected to plastic bending having
xviii CONTENTS

idealized stress-strain curve Fig. 22-8. — Plastic bending considering work


hardening according to idealized stress-strain curve of Fig. 22-13.
Examples — Load-carrying capacity of steel bars bent in plastic range of
strains.

23. Buckling After the Yield Point Is Exceeded 371

PART III
THE ELASTIC, THE VERY VISCOUS, THE IDEALLY PLASTIC SUBSTANCE
AND SOME OF THEIR GENERALIZATIONS
SPECIAL PROBLEMS OF THE IDEALLY PLASTIC SUBSTANCE
24. Synthesis of Small Elastic and Permanent Strains 379

<t>
Elastic, plastic and resultant strain tensor — Flow functions and

^.
Examples.

25. Isotropic Elastic Solid 388


The equations of the theory of elasticity in rectangular coordinates and in
vector form.
,

26. Steady Slow Flow of Very Viscous Substance 395


The equations of viscous substance. Plane problem. Equations in
a

vector form.

27. Ideally Plastic Substance 401

28. Further Simple and Composite Plastic and Viscous Solids 405
Constant maximum shearing stress — Generalized ideally plastic sub
stance — Theory of the strain hardening of metals —Steady stage of creep —
Plastico-viscous substance — The elastico-viscous and the firmo-viscous
solid — Plastic deformation considering the elasticity of the material.

29. Constrained Flow of Ideally Plastic Material. Stress-Tensor


Trails Resultant Strain Tensor 426

Constrained flow neglecting elasticity — Constrained flow considering elas


ticity — Example of case of combined tension and torsion — Flow when
components of stress-tensor remain constant —Same cases including elastic
strains — case in which the principal axes of the tensor of stress rotate —
A

Example by E. Reuss.

30. Theory of the Plastic Deformation of Thick-walled Cylinders.


Cylinder of Ideally Plastic Material 436

State of plane strain in ideally plastic cylinder — Ideally plastic cylinder


without axial extension — Radial and axial flow.

31. Flow in Metal Cylinders Considering Strain Hardening 448

Theory for small strains —Special cases of cylinders with closed ends — Plas
tic flow in moderately thick-walled cylinders — Cylinder with closed ends
yielding under general strain hardening law at finite strains — Instability
of equilibrium causing local bulging in plastically deforming cylinder.
CONTENTS xix

32. Partial Yielding in Cylinders -158

First yielding in cylinder under internal pressure — Comparison of three


cases — Partial yielding of ideally plastic cylinder in plane strain — The
partial yielding in cylinder having closed ends under an internal pressure —
Spreading of plastic zone around a cylindrical cavity in which the pressure
is increased.

33. Theory of the Plastic Deformation of Flat Rings ob Disks ....


Flat ring of ideally plastic material — Permanent expansion of a hole in a
472

steel plate — Infinite plastic disk in tension having a circular hole.

34. Distribution of Stress in Rotating Cylinders and Disks 482


Rotating cylinder under plane strain — Comparison with elastic rotating
cylinder — Rotating disk — Solid disk — Rotating disk with hole.

35. The Problem of Plastic Torsion. Experimental Representation of


Stress Distribution 490
Elastic torsion. Prandtl's soap-film analogy. The elastic stress function
for torsion — The plastic stress function of torsion. The sand-heap anal
ogy — Apparatus for experimental determination of stress distribution —
Yielding in bars having holes. Hollow rectangular cross section treated
by F. S. Shaw — Partial yielding in a twisted shaft of variable cross section.
Eddy's and Shaw's solution.

36. The Flow Layers in Twisted Bars of Mild Steel. Effect of Grooves
and Holes 512
Flow layers. Torsion tests with steel bars — Structure due to cooling and
fluidal structure — Longitudinal groove with semicircular cross section —
Cylindrical hole — Bars with longitudinal grooves — Instability of uniform
mode of yielding in a twisted round bar of mild steel — Disturbance by
flow layer analyzed by E. Reuss.

37. Plane Strain and Plane Stress. Theory of the Surfaces of Slip . . 527
Plane strain, e, = 0 —-Various ways for obtaining solutions — Equations

for the slip lines— Plastic mass pressed between two rough parallel plates —
Radial flow — Circles as envelopes of slip lines — Vortical flow in a plastic
mass — Radial distribution of stress in a wedge-shaped plastic space — The
equations of plane flow of an ideally plastic substance expressed in curvi
linear coordinates through the slip lines — Geometric properties of plane
fields of slip — The characteristics of the partial differential equations in
the theory of the plane flow of an ideally plastic substance are the lines of
slip — Displacement field under state of uniform stress — Characteristics of
partial second order hyperbolic differential equation of two independent
variables — The two, orthogonal families of the slip lines are the character
istics — Plastic states of plane stress — Conclusions — Three principal meth
ods summarized for the treatment of problems on flow of solids.

Index 661
LETTER SYMBOLS
x, y, z rectangular coordinates
<t>

r, <p, or polar coordinates


cylindrical coordinates
r,
z

time
t

a normal stress
t shearing stress
<rx, <r„, ff, components of normal stress in rectangular coordi
nates
' *»J ' V*1 ' ** components of shearing stress in rectangular
coordinates
ffl, (72, 03 principal stresses
Tl, Ti, U principal shearing stresses
a r. <r,, a. radial, tangential, axial normal stress
a,. yield stress in tension
Too or t„ octahedral shearing stress
t

a =
{<r

+ + mean stress
3

ff2 o-3)
:
i

hydrostatic pressure
V

unit strain (conventional strain)


i

unit shear
1

ti, «», «* unit strains in the directions of the x, y, axes


z

"Yxv, *Yw*? y*x unit shears


«r, (i, «• unit strains in radial, tangential, and axial
direction
•l, «!, «J principal strains
= + quadratic elongation
(1

«)2
X

a. normal and shearing stresses, referred to deformed


f

state of body
e),

= In natural strain and shear


+
(1

y
i

t" elastic and permanent part of unit strain


«',

a a

,y" elastic and permanent part of unit shear


E, y'

K
G,

modulus of elasticity, rigidity, bulk modulus


Poisson's ratio
V

Lame's constants of elasticity


X.
M

ri, or u, w components of displacement


v,
1

«, w components of velocity
V,

*. ¥ flow function for tension and for shear


xxii LETTER SYMBOLS

D.D. deviator of stress and of strain


ax, Oy, a. direction cosines
r»v variable vectors
a, b, c constant vectors
i. J, k unit vectors
a • b scalar product of vectors
a Xb vector (skew) product
*, T,n dyadics
I idem factor
V nabla operator
a2 si
A = 1 Laplacian operator in two dimensions
dx2 dy*
div v divergence of vector
rot v rotation of vector
grad gradient
.1/ moment of a couple, torque
6, <p angle of twist
P,Q single force
I moment of inertia of a plane section
A area of a plane figure
V volume
S area of a surface
P radius of curvature
b,h width (breadth) and height of a cross section
h thickness-of a plate or shell
y deflection of a beam
w deflection of a plate
V- coefficient of viscosity
p density
7 specific weight (weight per unit volume)
9 gravity acceleration
T absolute temperature #
a linear coefficient of thermal expansion
w work

Attention of the reader is called to "Letter Symbols for Mechanics of


Solid Bodies," ASA Z10.3, 1948, issued by the American Standards
Association and recommended for use by the four American engineering
societies, published by the American Society of Mechanical Engineers.
The symbols in the preceding essentially follow these recommended
standards.
Parti
DEFORMATION OF SOLIDS. ANALYSIS OF
STRESS AND STRAIN
CHAPTER 1

INTRODUCTION
In discussions with sensible professional men, I have not infrequently
encountered the opinion expressed that it would be wasting vain efforts
to develop a theory on which the strength of materials could be based
scientifically. Homogeneous bodies of materials — I was told — do not
exist, homogeneous states of stress are not encountered. It seems,
therefore, utterly impossible to deduce a law of nature from experience.
Since the existing irregularities furthermore are of such a nature that
they nearly completely obscure any orderly behaviour, it has little
interest to track down the half-blurred traces of such laws. Under
these circumstances nothing else remains than to make special tests
in every important case and to pay no heed to a physical interpretation
of the results. I had to admit in each case that nothing could be said
against this reasoning ; and yet for more than one hundred years, there
have been attempts again and again to establish order within the con
fusing abundance of the experiences. If one should succeed in finding
a few rules under which many experiences could be subordinated — of
course rules in which some confidence can be placed — no law of nature
would have been derived, but some means found for judging the prob
ability of new results of experience.
Mohr, Otto, Z. Ver. deut. Ing., p. 740, 1901.
The metallic elements in the periodic system of the chemical elements
have remarkable physical properties. They can be deformed per
manently by large amounts at temperatures considerably below their
melting temperature and at elevated temperatures. Single crystals of
the elements iron, aluminum, copper, and zinc and of other metals in the
form of cylinders or prisms at normal temperatures can be extended
permanently under tensile forces by surprisingly large amounts before
they break. The beautiful experiments, made a few years ago by various
investigators, in which metallic single crystals were permanently elon
gated by a gradual increase of the tensile forces to considerable values are
the purest manifestations of one of the outstanding mechanical proper
ties of metals. In the polycrystalline form the metallic materials of con
struction have additional remarkable properties. We might mention
here their ability to deform very little elastically under load as long as
the load does not exceed certain values. Concerning their ability to
deform permanently to a considerable extent under further increase of
load, we may include here the property that the polycrystalline metals
3
4 THEORY OF FLOW AND FRACTURE OF SOLIDS

can be formed cold or hot through forging, bending, hammering, pressing,


drawing, rolling, etc. Steels and ferrous and nonferrous alloys can be
hardened. In their hardened state, they yield to the permanent deforma
tion at much higher loads than in their soft condition.
At low temperatures which are considerably below the melting point
of a metal the instantaneous values of the forces during a loading period
usually determine the corresponding state of deformation in the body.
At higher temperatures which approach the melting point of a metal,
this is no longer true. A metal piece carrying a constant load over a long
time at a moderately elevated temperature slowly deforms in a continuous
way. This kind of slow permanent deformation is known as creep. For
predicting the total deformation evidently the time must be considered as
a new additional independent variable. This is true for all solids at
sufficiently high temperatures, for some in a more pronounced manner,
for others to a lesser degree. The rate at which the creep changes with
respect to the time at a given load for a given material increases very
rapidly with increasing temperatures.
Associated with the slow creep of metals at elevated temperature is
another general phenomenon in metals to which reference should be made.
If a spring of good steel is pulled out and the load released soon thereafter,
the spring contracts back to its original length even if this occurred at a
fairly high temperature. If the spring
is,

however, held in its stretched


condition over a long time at the high temperature so that the ends are
not allowed to move, the force exerted by the spring will gradually start to
diminish. After sufficient time has elapsed, the force may have com
pletely disappeared, and in this condition the spring may not show any
more tendency to contract. This phenomenon known as relaxation.
is

Cast metals, particularly the ones having a porous structure or con


taining soft inclusions such as the graphite particles in cast iron, are
seldom used under tensile forces. They may, however, withstand com
paratively great forces in compression without appreciable distortion.
Parts of machines have recently been produced from metallic powders
through pressing between dies and subsequent sintering in which sur
prisingly high strength values were claimed.
The properties of natural rocks are to some degree comparable with
those of artificial materials. The granites, basalts, etc., in other words,
igneous rocks which originally existed in the molten state before they
erupted and solidified, may be compared in their properties with the cast
metals. Such rocks have high compression strength. The same,
is a

although to a lesser degree, true of more compact sedimentary rocks


such as sand- or limestones. The latter in nature were formed in a manner
somewhat analogous to the process by which articles from metallic
INTRODUCTION 5

powders are produced. After loose particles of silt, sand, or shells


accumulated along the shores of former seas, their layers were consoli
dated through the weight of other layers which were deposited above
them. The pressure of the overlying strata cemented the loose particles
into the compact masses of sand- or limestones.
Igneous and sedimentary rocks under ordinary compression loads
deform very little before they break. It is of great interest to note,
however, that evidences disclosed in the distortion of the strata in the
earth crust and in the grain structure of certain rocks point to the fact
that many rocks in greater depth have been severely and permanently
deformed during former geologic epochs. Extremely small differences
between the vertical and the horizontal pressures in the deeply buried
rocks were sufficient to produce the observed folds in the strata because
of the very long times during which these forces acted. These conditions,
together with the elevated temperatures and the high average pressures
prevalent in great depths, made the flow of solid rocks possible. Thus
similar phenomena have occurred and are still at work on a large scale in
the upper earth layers which bear resemblance to well-known observa
tions in small-scale laboratory experiments when engineers investigate
the flow of metals at normal and at elevated temperatures. For example,
geologists have found that certain elongated inclusions in volcanic rocks
are oriented symmetrically around the necks of extinct volcanos. They
were oriented in the positions they would have assumed if they had been
suspended and carried in a semifluid mass flowing in an upward direction.
This occurred during the period of dying volcanic activity when the
magmatic masses started to solidify. Structural geologists have con
clusively demonstrated by numerous observations that entire portions
of mountains must have been carried
away, folded, and sometimes
pushed above layers of rocks much younger than those of which they
consisted. Huge portions of mountain chains of certain older strata
were found in positions in which they were resting on layers of much
younger geologic age.
The salt domes in the states situated near the Gulf of Mexico offer
another classical example found in nature for the flow of a solid rock,
namely, of rock salt. These large accumulations of salt under the ground
in hundreds of examples have been carefully surveyed by American
geophysicists because of their importance for locating the oils and deposits
nearest to the surface, which offer the most favorable conditions for
drilling for oil. Originally the salt was deposited in thin horizontal layers
from receding portions of the sea. Subsequently it was buried by sedi
ments. Since the density of salt is slightly lower than that of the over
lying rocks, buoyancy forces must have been present in the salt layer.
6 THEORY OF FLOW AND FRACTURE OF SOLIDS

There may also have been weaker portions in the overlying rock strata or
points around which their density was slightly lower than at other points.
This would have been sufficient to cause a slow convergent flow in the
mobile salt toward these centers because of the tendency of a lighter
material to escape from beneath a layer of a heavier material (a layer of
oil would do this if it were left under water).
Reference should finally be made to the group of modern "plastics"
such as nylon, rayon, bakelite, and other highly polymerized organic
materials, which have been developed recently by the chemical industry,
and to the rubberlike materials, many of which possess
most unusual and striking peculiarities from the point
of view of their mechanical behavior. These are due to
the filamentous or fibrous structure of their long chain
like molecules. Un vulcanized rubber has presented
some puzzling problems to engineers and physicists who
have tried to define some satisfactory means for the
measurement of the intensities of the large elongations
under which rubber in sheet or band form is capable of
being stretched and of the internal forces under which
it deforms. The length of a cord or band of very elastic
■ i ll, rubber may be increased twice to eight times. After
ill
Mi

removal of the stretching force may contract back to


it
1

II
I

nearly its original length. The volume during the


T

Fiq. 1-1. Con stretching remains practically unchanged. Although


striction forming
in such material the amounts of potential energy
a

on thin filament of
unoriented nylon which may thus be stored during stretching are con
after pulling.
siderable and practically recoverable upon unloading,
in contrast other types of rubber are valuable because of their property of
acting as dampers of oscillations. Rubberlike materials and plastics at
normal temperatures show time effects and creep.
The substance nylon product of du Pont) has a strange and interest
(a

ing property which we wish to describe very briefly. In the unstretched


condition, thin nylon filaments (or sheets) consist of long chainlike mole
cules which are oriented at random. Under tensile force which increases
a

from zero these filaments do not stretch much at first. When certain
a

critical load reached, however, has been found through X-ray analy
is

it

sis that the long molecule chains suddenly rearrange themselves so that
they become parallel to the axis of the filament. At this instant sharp
a

constriction forms in the profile of the filament (Figs. 1-1 and 1-2).
A

moment later the contracted portion seen to increase its length and
is

two "wave fronts" begin to move in opposite directions along the speci
men congruent in their shape to the halves of the necked portion. While
INTRODUCTION 7
8 THEORY OF FLOW AND FRACTURE OF SOLIDS

all this happens, the contracted portion becomes longer, the load remains
perfectly constant, and evidently new portions yield in which the mole
cules rearrange themselves. After the entire fiber has contracted, the
load starts again to increase considerably. We shall later show that this
process of stretching of a nylon filament is analogous to the well-known
manner in which a bar of mild steel elongates. This metal has the charac
teristic property of possessing a well-defined "yield" point, at which it
starts to stretch permanently. Although the processes in the molecular
or atomic structures causing this sudden "yielding" in these two mate
rials are completely different, it is remarkable to note that the modes in
which nylon and mild steel visibly stretch are analogous.
Certain soft high polymers (plastics) show another striking phenome
non. A short cylinder of such material may be compressed between
parallel pressure plates until its height is reduced by one-third or a similar
amount. Simultaneously, the cylinder is seen to expand laterally and to
take the characteristic barrel-shaped form of a compression specimen of
a ductile metal. If the load is suddenly released, this deformation
appears as a "permanent" one. If one waits long enough, however, the
cylinder slowly "recovers" from its "permanent" deformation until its
original shape is practically restored. Such partial and small "recovery"
from "permanent" deformations has even been noticed when the strong
est alloys of metals are tested at high temperatures. Metal bars, for
example, which were tested under a constant tensile load over an extended
period during which they deformed continuously through creep may
contract quite slowly to some extent over many months after having
been quickly unloaded and thus "recover" a portion of their permanent
extension.1
In view of the confusing complexity of the properties of solids with
reference to the phenomena of flow and of rupture it appears to be a hope
less task to record in detail in the- space available in one or two volumes
some of the most important results of experiments which have been
carried out in laboratories or to present these in a system which would
classify the solids, for example, according to their chemical composition
or their atomic or molecular structure. Fortunately, this seems to a
certain degree unnecessary for a reason which will become apparent.
When the innumerable observations gathered from mechanical tests for
Among the newly developed silicones there are soft grades which will flow under
1

their own weight. A ball which may be quite easily deformed through the pressure
exerted by both hands will soon start to spread under its own weight if left on a table.
The same ball when dropped from a height of several feet on the floor will jump back
and disclose a high grade of elasticity. Under very rapid straining this material
behaves nearly like a perfectly elastic material; under small stresses continuously
acting it deforms quite slowly permanently and flows like a viscous fluid.
INTRODUCTION 9

various materials are compared, it is soon discovered that substances


seemingly very different from a purely mechanical point of view fre
quently behave in a similar manner. In other words, if the right vari
ables which must serve to measure certain properties are defined, many
observable facts and phenomena can be expressed by identical relations.
This volume aims at the task of reviewing these common trends in the
behavior of solids by introducing a relatively small number of ideal sub
stances for which the properties are clearly defined and by analyzing
subsequently their behavior in special cases which are important in the
technical or physical applications. This is naturally done in the hope
that certain general results derived from theory may represent essentially
the actual behavior of certain much used materials or, if this should not
quite be the case, at least may offer quantitative information relative to
the forces required for deforming bodies or to the permanent change in
shape of bars, plates, cylinders, tubes, etc. This is of great value to
many industries, particularly to those forming the ductile metals through
rolling, drawing, extruding, etc., even although the conditions of manu
facturing admittedly are frequently quite complex in reality. In this
volume considerable attention will be devoted to a discussion of the
conditions under which certain classes of solids start to deform per
manently, or to what have been called "the theories of strength," and in
this volume the corresponding conditions under which solids break
will be reported. One type of failure through slip is of fundamental
significance through the entire theories of the permanent deformations
of solids. The beautiful geometric properties of these surfaces of slip
which have recently been discovered will be described in detail, some of
them constructed mathematically, and their application also to the
equilibriums of the forces in loose powdery substances of sand and earth
discussed.
CHAPTER 2

THE SOLID AND FLUID STATE OF MATTER.


ELASTIC, VISCOUS, AND PLASTIC SUBSTANCES
And swiftly as we see wines flow through a strainer,
Yet sluggish olive-oil will take its time,
Undoubtedly because it is composed
Of elements either larger or more hooked
And mutually entangled: and thus it is
These particles cannot become so quickly
Detached from one another, and percolate
Each singly through their several openings.

Lucretius (first century B.C.), translated by


R. C. Trevelyan, "Of Nature of Things," p. 57,
Cambridge University Press, London, 1937.

In the study of the mechanical properties of materials from a rational


viewpoint the simplest possible forms of the assumptions have proved to
be valuable. From a mechanical standpoint it is useful to ascribe certain
ideal properties to materials. For example, both the motion of material
elements and the states of equilibrium existing in them may be studied
from a common standpoint, if we attribute to a substance certain definite
simple properties. Thus we are led to such definitions as those of uni
formly distributed masses, of perfect fluids and gases, of viscous sub
stances, of isotropic elastic solids, etc.
In physics, three states of aggregation of matter are distinguished: the
solid, the liquid, and the gaseous state. In order to avoid difficulties and
obvious contradictions, however, it should be noted that this classification
is not essential and is fundamentally unnecessary in many cases and in
others is unsatisfactory.1
Bodies in the solid state usually maintain their shape under their own
weight, while fluids are characterized by the easy mobility of their small
elementary parts. In mechanics this has led to the definition of a "per
fect fluid" in the interior of which at rest the pressure in a point must
always be directed perpendicularly to any infinitesimal section which

1 Thus, as long as the velocity of flow of air is much smaller than the velocity of
sound, the flow of air around a solid body may be similar to the flow of water around
the same object. For moderate velocities the flow of a compressible gas such as air
may follow the same laws as the motion of an incompressible fluid such as water.
10

^
ELASTIC, VISCOUS, AND PLASTIC SUBSTANCES 11

may be assumed in any arbitrary direction in the fluid passing through


the point. (A component of force which acted in a tangential direction
through a section would displace the adjoining fluid particles in the same
way that a wooden board which floats on the surface of water would start
to move if the smallest force acted on the board in a direction parallel
to the surface.) The elements of a solid on the contrary can also trans
mit, in addition to normal components of the forces, tangential com
ponents of considerable magnitude through any section.
Solid materials are said to possess a definite melting point at atmos
pheric pressures. Melting of solids is usually accompanied by a sudden
change of the volume. (This is usually an increase in volume, with a few
exceptions, the best known of which is ice. The volume of solid water at
atmospheric pressure decreases upon melting.) Amorphous materials,
however, may be brought from the solid into the fluid state without a
sudden change in volume, and no definite temperature can be observed at
which they melt. If heated quite gradually, they pass over a range of
the temperatures from the "solid" to the "fluid" state. Examples are
most glasses, asphalts, plastics, etc., which gradually soften if heated
and harden if cooled.
Although almost all fluids at rest behave as perfect fluids, the presence
of tangential forces can be observed in
these substances during motion. A heavy
lubricating oil exerts a measurable though
small resistance against its motion if it is
brought between two parallel plates closely
spaced and one plate is moved relative to
the second one in a direction parallel to the plates. It will be found that
the force per unit of area of the contact surface t measured in pounds
per square inch is proportional to the relative sliding velocity u and
inversely proportional to the distance a of the two plates. Thus a fluid
exerts an internal resistance against this motion. This is expressed by

' -? <™

or, in a more general manner, after introducing rectangular coordinates


x in the direction of motion and y in the direction perpendicular to the
plates, also by

(2-2)

The material constant m is called the coefficient of viscosity, and such a


12 THEORY OF FLOW AND FRACTURE OF SOLIDS

substance a viscous one, this property being the viscosity of a substance.1


When glasses are heated slightly above their "softening" temperatures,
they behave in this manner as viscous fluids. At atmospheric tempera
tures, they are considered to be in the "solid" state. Since there is no
definite temperature at which a glass melts, one would expect that their
viscosity would also become apparent at temperatures below their range
of softening. We must thus note that (1) a "solid" glass at lower tem
peratures must behave like a "viscous fluid," (2) glass in its solid state
would be apt to deform permanently, and (3) since the behavior of solids
interests us, in general, over a wide range of the temperatures and since
the viscosity of fluids is known to change and to increase with decreasing
temperatures, of glassy materials should vary considerably
the viscosity
and increase to very large values compared with those slightly above the
softening temperatures if the temperatures become low. All these con
clusions have been fully verified by experiments.

Fig. 2-2. Fig. 2-3.


Figs. 2-2 and 2-3. Creep of candles. (The deformation occurred during several exception
ally hot summer days.)

Gradual permanent changes in shape at relatively low temperatures, at


which certain amorphous materials, such as bituminous solids, asphalts,
and tars, are usually considered as solids, may be observed under the
action of quite small forces, such as, for example, the weights of bodies.
It is only necessary to allow sufficient time for the action to take place.
Hard asphalt, which may shatter into pieces through a hard blow, may
flow slowly out of a crack from a leaky barrel. If a steel ball is placed on
1 Observations of the flow of fluids around solid
bodies and general considerations
of the mechanisms of viscous (and of turbulent) flow in aero- and hydrodynamics
have offered convincing proof that substances such as water or air, which in certain
phenomena are considered to behave as "perfect fluids," near the walls of solid
bodies must transmit tangential forces (drag). Thus we note a case in which the
concept of a certain ideal substance must be abandoned: while water or air in motion
may be treated as perfect fluids at larger distances from a solid body, in the "boundary
layer" near it they behave like "viscous" (or "turbulent") fluids.
ELASTIC, VISCOUS, AND PLASTIC SUBSTANCES 13

the free surface of solid asphalt filling a container in warm weather it will
in time sink below it and after sufficient time will be found at the bottom
of the tank. A bar of sealing wax, supported horizontally at both ends,
may bend under its weight after a time. '
Moreover, the crystalline materials may deform permanently at tem
peratures considerably below their melting point. A number of examples
were cited in the preceding chapter. Among the polycrystalline mate
rials the formable metals take a prominent position through their property
to. deform permanently under tensile forces of a sufficient magnitude at
low temperatures. Laboratory experiments have proved that other
crystalline materials such as, for example, the brittle natural rocks
(marble, sandstone) can be deformed continuously to considerable
amounts permanently in short time tests. To achieve this, however,
high compression forces are required, which must be applied in different
directions relative to the specimens with different intensities.
On the basis of these few observations, to which countless others could
be added, it may be concluded that all solid bodies under appropriate
conditions may be brought into a state in which they change their shapes
continuously without failing through fracture. It is an experimentally
established fact that even the strongest metals, such as the high-strength
alloys of iron, as well as nonmetallic materials with a polycrystalline
grain structure, such as minerals and brittle rocks, under certain definite
mechanical conditions may be brought into this "plastic" state, in which
permanent deformation may occur without fracture. Solid bodies or
materials in this condition are said to "yield," to "flow," or to deform
1 Sometimes the opinion is expressed that glass tubes, if stored at room temper
ature and freely supported at their ends, would bend permanently under their own
weight. Lord Rayleigh (Nature, Mar. 1, 1930, p. 31 1) found that a glass rod 4.9 mm
in diameter and about 1 m long, supported freely at its ends and loaded by 300 g at
the middle, which produced an elastic deflection of about 2.8 cm, did not show an
appreciable permanent increase in its deflection after a period of 7 years. Apparently
the viscosity coefficient m of glass at room temperature is so large that it would require
a much longer time to produce measurable effects at this low temperature and com
paratively low stress. C. D. Spencer (Nature May 10, 1930, p. 707) found, however,
that a glass tube of 110 cm length, 1 cm diameter, and 1 mm in wall thickness on sup
ports 1 m apart and loaded by 885 g at the center (a load which was little less than
the force necessary to break the tube in bending) showed a permanent deflection of
9 mm after the lapse of 6 years of loading.
Wooden beams in old frame buildings which have carried their own weights or
additional loads over several centuries frequently show considerable sagging. Stones
or slabs in tombs in old cemeteries which have been supported in horizontal position
at their ends have been seen in sagged or bent condition (E. C. Bingham). In similar
cases additional circumstances (exposure to severely changing weather conditions,
intermittent freezing of water in the cracks with subsequent thawing, etc.) might have
contributed to the distortion and been at work producing the observed effects.
11 THEORY OF FLOW AND FRACTURE OF SOLIDS

plastically if the permanent changes in shape are sufficiently great to be


perceptible.
Another universal property of crystalline and amorphous solids is their
elasticity. A relatively small but easily measurable portion of the total
distortion of solid bodies under load is said to be of an elastic nature. The
crystals of hard minerals (quartz, diamond) are examples of bodies which
may illustrate this fundamental property in its purest manifestation;
under forces homogeneously distributed, they deform by very small
corresponding amounts, which depend solely on their instantaneous
values. On release of the forces, these small amounts of distortion dis
appear perfectly. The changes in shape also depend on the direction

TANGENTIAL
FORCE FORCE

YIELDPOINT
Y-

DEFORMATION RATE OF SHEAR DEFORMATION


Fig. 2-4.Elastic Fig. 2-5. Viscous Fig. 2-6. Plastic
deformation. deformation. deformation.

relative to the axes of the crystals along which these are stressed and on
the symmetry properties of the crystals. Furthermore, these changes
are proportional to the applied forces. In crystal physics, such dis
tortions are named as anisotropic elastic distortions. The directional
effects of them cancel out in the polycrystalline solids because of the
random orientation of the crystal axes in the small crystallites. This is
more the case in amorphous solids in which the particles are supposed to
be submicroscopically small. Relative to the small reversible distortions
ordinary solids have isotropic elasticity.
Instead of distinguishing in general the solid from the fluid state of
matter and for the sake of avoiding some of the afore-mentioned obvious
contradictions in its postulated behavior, it is better to speak of certain
idealized substances the mechanical properties of which are clearly
described. This leads us to define three idealized substances first which
we have already characterized somewhat by the modes according to
which they deform. They may be characterized by means of the sym
bolic graphs shown in Figs. 2-4 to 2-6.
To begin with the elastic substance, since the forces are proportional to
the distortions (expressed in some convenient manner to be defined later)
an elastic substance is represented through a straight line as shown in
ELASTIC, VISCOUS, AND PLASTIC SUBSTANCES 15

Fig. 2-4. Similarly a viscous substance may be represented by plotting


the tangential force per unit of area t [Eqs. (2-1) and (2-2)] in function of
the quantity u/a, which expresses the relative slip per unit of time in two
parallel planes gliding with the substance one unit of length apart.
"
u/a or du/dy is "the rate of shear (Fig. 2-5). Finally we may construct
a diagram characterizing a perfectly plastic substance by assuming that the
force per unit of area causing an extension or contraction must reach a
certain finite value for the initiation of the permanent distortion. When
this characteristic load is reached, the subtance starts to deform per
manently or to "yield" and continues so at this load. Assuming that the
changes in shape remain small, this case is then represented by drawing a
horizontal straight line in the diagram in which the permanent exten
sions are plotted as abscissas and the forces per unit of area as ordinates
(Fig. 2-6).
A few important materials deform almost according to one or the other
of the three pure modes we just described (at least as long as the forces or
the extensions do not exceed certain given limiting values). In general,
it will be found, however, that the behavior of materials is more com
plicated and that (1) in solids distortions of an elastic, viscous, or plastic
nature may occur simultaneously. Combinations of two types of these
three kinds of distortions are known to occur frequently. Those cases
are of particular interest in which the total deformation appears as the
sum of an elastic and a viscous or a plastic portion. (2) The forces
causing the permanent portion of an extension or compression may
appear as functions either of the permanent deformations themselves or
of their rates with respect to the time or of both. Considerable attention
will be devoted to the treatment
of these more general cases of the flow of
solids in this volume after the theories of the simpler substances have
been discussed.
The flow of certain dense suspensions of small particles of a solid in a
viscous fluid has been treated by E. C. Bingham.1 He assumed that the
force must reach a finite or "yield" value before the flow can start.
Examples for such plaslico-viscous substances according to him seem to be
certain paints.
To explain the phenomenon of relaxation in its purest manifestation in
solids it suffices after J. C. Maxwell2 to assume for the total extension that

1 "Fluidity and Plasticity," pp. 217, 219, McGraw-Hill Book


Company, Inc., New
York, 1922, 440 pp. Also Reiner, M. "Ten Lectures on Theoretical Rheology,"
p. 133, Rubin Mass, Jerusalem, 1943, 163 pp.
*Phil. Mag., vol. 35, p. 134, 1868. Also Jeffreys, Harold, "The Earth, Its
Origin, History and Physical Constitution," 2d ed., p. 265, The Macmillan
Company, New York, 1929. Cf. also Chap. 28, Sec. 6.
16 THEORY OF FLOW AND FRACTURE OF SOLIDS

it consists of the algebraic sum of an elastic anda purely viscous portion.


More complicated cases of relaxation may be treated also. These exam
ples, together with cases which have previously been mentioned, may be
based on theoretical analysis showing readers that, in spite of the admit
tedly complex nature of the phenomena associated with the permanent
deformation of solids, an analysis may offer valuable information concern
ing observable facts. The properties of some materials may be illustrated
by certain simple mechanical models which have been proposed by Dutch
and other scientists and may be helpful also for explaining idealized
examples of relaxation, recovery, and similar phenomena. 1
1 "First and Second Reports on Viscosity and Plasticity," prepared by the Com

mittee for the Study of Viscosity of the Academy of Sciences, Amsterdam, 1935 and
1938 (first report 256 pp., second report 287 pp.). The mechanical models are
described in the first volume in an article by T. M. Burgers, pp. 5-64.
CHAPTER 3

ELASTIC AND PERMANENT DEFORMATION

3-1. Conventional Stress-Strain Curve. If a bar of ductile metal with


a constant cross section is loaded in a tensile machine, it will lengthen
uniformly under the action of an applied load. In order to record the
changes of length suppose that two lines were marked crosswise on the bar
and that the distance between them is designated by Under the ten

l0.
sion load P, the length will increase to length If the extension
l0

I.
Al = — plotted against the corresponding load in a system of

P
is
lo
I

rectangular coordinates (Fig. 3-1), continuous curve will be obtained.


a

kLOAD

TENSION

Fio. 3-1. Load-extension curve.

Instead of tracing this curve, a standard practice to divide the exten


it
is

sions Al by the original length of the bar and the load P by the area A0
lo

of the original cross section and to plot the curve of P/A0 in function of
Al/l0. The extension Al divided by the original length called the
is
l0

unit elongation or strain and the load divided by the area A0 the unit
P

stress in the bar. The former will be designated by = Al/l0 and the
e

latter by <r = P/Aq. The curve <r /(e)= called the conventional stress-
is

strain curve (diagram) of the metal.


the temperature at which the metal tested considerably below
If

is

is

its melting temperature neither the rate at which the stresses a increase
nor the rate at which the elongation increases during the test has
a
e

marked effect on the shape of the stress-strain curve. While the load
P

or the stress a increases monotonously we may therefore assume that the


stress a function <r = /(e) of the strain e.
is
a

17
18 THEORY OF FLOW AND FRACTURE OF SOLIDS

If a wrought-iron or mild steel bar is loaded


in tension, an initial
stress-strain /(«) is observed such as is shown in its first portion
curve a =
in two examples in Figs. 3-2 and 3-3. The stress a at first increases in
exact proportion of the increase in strain *. At a certain stress <fX (point
A in Fig. 3-2) or <r0 (point B in Fig. 3-3) permanent deformation suddenly
begins to appear. Along the steep straight
part of the curve as the load increases the
extension is essentially elastic. If the bar
is unloaded before the curve reaches the
point A or starts to bend toward B, re
spectively, no permanent set will be observ
able. At the time the stresses o\ or <m,
STRAItF respectively, are reached sudden breaks
Fia. 3-2. Fia. 3-3. occur in the stress-strain diagrams and the
Fios. 3-2 and 3-3. Stress-strain strains increase greatly. The tensile stress
curves. Transition from elastic
to plastic stage (mild steel). ci (or cro) is called the yield stress. It indi
cates the start of full yielding through some
cross section of the bar. The transition to the horizontal portion of the
diagram either can begin with a sharp peak and subsequent rapid decrease
in the load as shown in Fig. 3-2; or it may be gradual or the steep straight
portion of the diagram may turn in a very sharp bend in the horizontal
branch, as shown in Fig. 3-3.

Fio. 3-4. Hard steel. Fio. 3-5. Soft copper. Fio. 3-6. Mild steel.
Fias. 3-4, 3-5, and 3-6. Stress-strain curves.

Higher carbon steels usually show a gradual transition from the elastic
to the plastic stage (Fig. 3-4). A yield point may not become observable
in such a case. The point P at which the curve starts to deviate from the
straight line has been called the proportional limit. It indicates also the
stress above which the permanent strains (permanent sets after reloading)
become appreciable.
ELASTIC AND PERMANENT DEFORMATION 19

Besides the examples indicated in the case of very soft metals a curve occurs such
as that represented in Fig. 3-5. If a bar of soft annealed copper or aluminum is loaded
in tension, the stress-strain curve <r = /(«) begins gradually to bend and the permanent
deformations start to develop at the smallest stresses. Since the slope of the steeply
inclined straight portions of the former stress-strain curves, or a It in the theory of
elasticity, is defined as the modulus of elasticity E of the metal, we may note that the
slope dtr/dt at which the stress-strain curve in the last example (Fig. 3-5) is inclined
toward the t axis at the origin 0 must measure this important physical constant in the
case of a soft copper or aluminum, for example. The tangent line OB at point 0 to
the curve indicates this slope in Fig. 3-5. The total strain t corresponding to a point
A and to a stress <r (Fig. 3-5) on the stress-strain curve is the sum of an elastic strain
t = a IE and a permanent strain t". These two strains can be measured off along the
horizontal line laid through A in the distances CB = t' and BA = t".
If the hardness of the metal is measured by means of one of the usual hardness test
ing methods, e.g., through Brinell tests, after the metal has been stretched permanently
by different amounts it will be found that its hardness has increased continuously
somewhat in proportion to the ordinates of the stress-strain curve. Metals are said
to have been cold worked, or strain hardened, through such stretching.
If a bar is unloaded after it is stretched to the stress a*, corresponding to the point D
in Fig. 3-5, the curve on unloading will be as shown in the figure by the branch D2E,
which is different from the loading curve 0\D. Suppose that the bar is again reloaded.
Then a new branch EZF results, which is different from the last unloading curve 2.
The unloading and loading curves 2 and 3 (Fig. 3-5) usually differ little from each
other (actually they form a very narrow loop). The loading curve 3 passes very near
to the point D from which the unloading branch 2 dropped down and line 3 bends at a
stress which differs little from o-* sharply in a direction which coincides practically
with the tangent to the original portion of the stress-strain curve at point D. In first
approximation, one may substitute a straight line DE for the narrow loop circum
scribed by the unloading and reloading branches 2 and 3. The slope of this line is
found to be practically the same as that of the line OB. Thus we note that the
modulus of elasticity E remains practically unchanged after a metal is cold worked.
We note also that, although a soft metal such as copper or aluminum has no yield
point in the annealed condition, it does appear after the metal has been cold worked
by stretching. For example, if the metal was stretched to the stress <r* (correspond
ing to the point D in Fig. 3-5), thereafter unloaded, and subsequently reloaded in ten
sion, a yield point and a new yield stress a* will be observed at which it continues to
stretch permanently.
The transition between a second loading curve such as 3-3 (Fig. 3-6) and the con
tinuation of the initial curve 1-1 in the case of soft iron or mild steel may be of various
shapes. If the tensile test is interrupted for only a short time by unloading and the
bar is immediately reloaded again, the curve bends over according to the line 3-3 in
Fig. 3-6. If, however, after an initial stretching and subsequent unloading (4-4) a
long time has passed (several hours or days) before the bar is again reloaded (5-5),
the yield point may appear raised, as Bauschinger and Ludwik have established, in the
manner indicated by the sharp peak of the reloading curve 5-5 in Fig. 3-6. This
transition is such that the sharp peak with the subsequent steep drop in load which
occurred at the original yield point Fas shown in Fig. 3-6 is repeated or even surpassed.

To illustrate further the behavior of a ductile metal at normal tempera


tures several stress-strain curves which were observed by testing an
20 THEORY OF FLOW AND FRACTURE OF SOLIDS

electrolytic copper are reproduced in Fig. 3-7. A series of cylindrical


annealed test pieces were prepared from the same copper. These were
stretched by different amounts in tension and their stress-strain curves
recorded up to the loads at which the tests were interrupted. As one
would expect for homogeneous material, these curves practically coincide
with the curve that was obtained for the bar that was stretched to the
maximum amount. From each of these prestretched bars, smaller
cylinders were machined and were subsequently tested in compression.

STRESS STRESS
4,,1 73lb/ m2 <r,— —4 .30* lO'LB, •in'
& „. Oh
-

1, 2, 3, 4, 1, .
0 —*-% ELONGATION %ELC NO SAVON ( I3X),

Jj
I - O
r 3 D 1 4 E 5 i S 1
'

44

u
a

I
h

I
i
\_B

,j— —
x

E — S
F
j
3S
\

G
T

- 4, 1,s
1

Fio. 3-7. Copper. Stress-strain diagrams of tension and subsequent compression tests
on copper at normal temperature (tests by E. A. Davis, B. A. Rose, and author, 1934).
Curves: A. Tensile test on soft annealed copper. B. Compression test on copper pre
stretched 1.5 per cent. C. Compression test on copper prestretched 2.5 per cent. D.
Compression test on copper prestretched 3.5 per cent. E. Compression test on copper
prestretched 5., per cent. F. Compression test on copper prestretched 1.,31 to 1.0 in.
diameter. G. Compression test on copper prestretched 1.,625 to 1.0 in. diameter. Upper
left curves: Soft annealed copper, stress-strain diagram at normal temperature. Note
that starting point of curve is not in correct relative position to other curves.
G

The stress-strain curves for these compression tests are reproduced in


the lower part of Fig. 3-7. This family of curves should then describe
how very ductile metal, such as copper, behaves after first strain
it
is
a

hardened in tension and subsequently deformed by compression. It will


be noted that all compression branches are sharply curved. Although
copper prestretched in tension, unloaded and again cold worked in ten
sion, has well-defined elastic range and sharp yield point as has
a

already been mentioned, will be seen now that after the stress
it

is

reversed
from tension to compression both the elastic range and a yield point for the
reversed direction straining have completely disappeared.1 Analogous
of

Certain of these observations were first pointed out by Bauschinger (Mitt. mech.
1

Tech. Laboratorium, Munich, vol. 13, 1886). The disappearance of the elastic range
and of the yield point after the direction of the applied stress reversed named after
is

is

him, "the Bauschinger effect." In the dissertations of W. Bader and W. Lode,


ELASTIC AND PERMANENT DEFORMATION 21

observations have been made by twisting a round bar permanently in one


direction and subsequently twisting it in the opposite direction.
3-2. Yield Stress of Polycrystalline Solids Depends Also on Rate of
Strain. In the ordinary type of tensile testing machine one head of the
test bar is connected to the force measuring system and displaced very
little during a test, while the second head connected to the driving
mechanism (screw drive or hydraulic piston) is usually moved with a
uniform velocity. The relative motion between the two heads of the
test bar determines the velocity with which it stretches in the cylindrical
gauge length in a tensile test or the rate of strain dt/dt with which the
permanent strain t increases with the time t. If the heads are moved
with a uniform velocity the rate of strain will remain constant in every
point of the bar as long as the gauge length elongates uniformly, i.e.,
before necking starts. Observations with the lower melting metals such
as tin, lead, and zinc have shown that the tensile stress-strain curves of
these metals have higher ordinates when they are pulled in shorter times,
i.e., under higher rates of strain than when they are stretched slowly.
Probably the first observer who established a mathematical relation for
the dependence of the stresses under which some of these metals yield in
tension tests with the rates of strain was P. Ludwik,1 who found that,
if the loads were compared at some given permanent strain t with the
rates v = dt/dt with which the bars were pulled, a function

<r = <ri + <70 log, —i (v > v0)

where <r0, <n, and are constants would express the relation between the
v0

stress a and the rate of strain v. By comparing the values of the stress a
along a line t = const such as ABC in Fig. 3-8 in the stress-strain curves
which are recorded for constant rates of strain v = dt/dt = const, thus a
may be observed as a function of v. It would then follow (at least in the
neighborhood of the given ordinate t = const) if the logarithmic speed

University of Gottingen, 1927 and 1928) further observations of a similar nature were
described. Curved transitions from the elastic to the plastic branch of stress-strain
curves were particularly observed after a slight rotation of the principal stresses
occurred in the strained bodies.
1 "Elemente der technologischen Mechanik," pp. 44—47, Verlag Julius Springer,
Berlin, 1909 (tests with tin wires). See also Cassebaum, Ann. Physik, 4th ser.,
vol. 34, p. 106, 1911 (tests with soft iron). Deutler, Dissertation, Gottingen, 1932;
Pbandtl. L., Gedanken Modell zur kinctischen Theorie der festen Koerper, Z. angew.
Math. Mechanik, 1928, p. 85; Meyer, Eugen, Forschungsarb. Ver. deut. Ing., No. 295,
pp. 62-73, 1927; Koerber, H. Stobp, and It. S. Sack, Mitt. Kaiser-Wilhelm-Inst.
Eisenforsch. DOsseldorf, vol. 4, p. 11, 1922; vol. 7, p. 81, 1925; Siebel, E., and A. Pomp,
vol. 10, p. 63, 1928; Brinkmann, H., Dissertation, Hannover, 1933.
22 THEORY OF FLOW AND FRACTURE OF SOLIDS

law were valid that if the rates of strain v = dt/dt are increased in a
geometric progression the corresponding yield stress a must increase in
an arithmetic progression. This has been shown on the right side of
Fig. 3-8 in a "rate diagram," in which the stress is plotted on semilog-
arithmic plot as a function of the logarithms of the rates of strain v as
a straight line.
When some of the ductile metals which have high melting points are
pulled, not much of this effect can be noticed at normal temperature
because of the comparatively narrow range within which the commer
cially available testing machines allow the plastic rates of strain to change.
Thus if we consider only such lengths of time as are necessary for carrying

'°^sTc.
STRAIN STRAIN RATE v —*-
Fio. 3-8. Influence of rate of strain on yield stress.

out the usual tensile test on some of the latter metals, we shall come to
the conclusion that the stresses at which yielding at normal temperatures
progresses depend but slightly on the speed of deformation. Metals can
be deformed fast through impact. By using a high-speed tensile machine
having a flywheel (Fig. 3-9) which is rotated at 300 to 1,000 rpm through
which, in the engaged position of movable horns, a bar can be broken in
small fractions of a second (J-Jjoo to 1/1,000 sec) and by recording the
stress-strain curves electrically on a cathode-ray oscillograph, the effect
of strain rates one million times faster than the rates available under
ordinary testing conditions can be observed. Stress-strain curves
recorded over this wide range of v = dt/dt are also found to be displaced
relative to each other within a band in the a, t plane (Fig. 3-8) for the
metals having higher melting temperatures, such as copper or mild steel.
The dependence of the true stresses (referred to the actual areas of the
cross section) at the ultimate stress ow, in copper is illustrated by the
curves in Fig. 3-10 in which the abscissas are proportional to the log
arithms of the rates of strain and the ordinates to the true stresses at the
ultimate stress o-m„. Each curve was obtained from a number of tensile
ELASTIC AND PERMANENT DEFORMATION 23

tests run at various rates of stretching at the temperature given above the
curve. The highest rates of strain were of the order of dt/dt = 1,000
(1/sec) corresponding to extending a bar by a strain equal to 1 (or by 100
per cent) in 1/1,000 sec. Figure 3-10 characterizes perhaps the general
trend of the stress-rate-of -strain function at normal and at elevated
temperatures for the polycrystalline (nonaging) ductile metals. It is
perhaps worth noting that at the highest temperature of 1000°C at which

Fig. 3-9. High-speed tensile machine built at the Westinghouse Research Laboratories
in 1939. The flywheel driven by an electric motor carries two retractable horns on both
sides which are moved by means of two steel springs during the rotation into their oper
ating position when a trigger mechanism releases them by means of a solenoid. The test
bar is mounted between the two steel columns in vertical position. On the right a small
induction coil can be seen which is slipped around the specimen for heating it by high-
frequency currents. The load and the extension of the bar are recorded by photoelectric
cells.

some of these tests were run, which was 83°C below the melting point of
copper, at the highest rates of strain equal to 1,000 (1/sec) in order to
overcome the internal resistance against a plastic deformation in copper,
a stress must be exerted which is one-third of the force required to stretch
this metal at normal temperature. At room temperature the curve
approaches a straight line over several cycles of the logarithmic scale
and thus approximately conforms with the equation given above.1
From these observations we must furthermore conclude that a definite
time is necessary to bring about the interruption of flow. During this
time the metal yields to some extent. Flow cannot be stopped instan-
1 Manjoine, M. J., and author, High Speed Tension Tests at Elevated Temper
atures, Part I, Proc. ASTM, vol. 40, 1940; Parts II and III, J.
Applied Mechanics,
June, 1941. Also Manjoine, Influence of Iiate of Strain and Temperature on Yield
Stresses of Mild Steel, J. Applied Mechanics, December, 1944.
24 THEORY OF FLOW AND FRACTURE OF SOLIDS

taneously after a certain stress has been reached and this stress is main
tained at a constant value (as is required for making an observation
during a tensile test) for a certain amount of time. Plastic strains
during this time increase by small amounts under decreasing rates. All
these phenomena are comprised under the heading of "afterflow." This
must cause an unloading line (such as D2E in Fig. 3-5) to become curved
in its upper portion, similarly a loading line (such as E3D in Fig. 3-5)

O~3 I0"4 IO~3 10~' 10~' I 10 O' I03


RATE OF STRAIN PER SECOND
Fio. 3-10. Stress-rate-of-strain function for pure copper tested in tension at seven
temperatures.

to bend in the opposite direction just before the bar starts to stretch
again plastically after a reloading. Pure "afterflow" cannot cause a
closed loop between these two branches of curves; only an inverted peak
(in the point E) should become observable.
3-3. Creep. At moderately high temperatures in solids, a continuous
deformation is observed under a constant load. An alloy-steel bar
heated to 500°C under a constant tensile force of sufficient magnitude
deforms permanently slowly and is said to creep. In engineering lab
oratories such "long-time" creep tests in tension are usually performed
over several months. Creep of metals at moderately high temperatures
from a mechanical standpoint is related to the phenomenon of viscosity
of amorphous solids mentionedin Chap. 2, although the laws are con
siderably different for these two groups of solids which express the depend
ence of the rates of creep on the stresses. Usually the strain due to creep
ELASTIC AND PERMANENT DEFORMATION 25

e is plotted as a function of the time t while the load is maintained con


stant. An example for such creep curves « = /(<) is represented for
several loads <r in Fig. 3-11 for an oxygen-free copper after tests by G. H.

Heiser (Westinghouse Research Laboratories) at 200°C. The total creep


strain t consists of an elastic strain «' = a/E and of the permanent portion
t". At comparatively small stresses one can usually distinguish three

0 50 IOO ISO 20O 250 300 350 400 450 500 550 600 650 TOO 750 800
TIME -HOURS
Fio. 3-11. The creep of oxygen-free copper at 200°C. The copper was first prestretched
by 8 per cent at room temperature. (After tests made by G. H. Heiser in 1939 in Pittsburgh.)

distinct portions in a creep curve. The first portion is curved and is


called the period of primary creep. During the secondary creep period
the creep curve straightens out almost perfectly. A strong metal may
creep for years at a constant rate. A good example of the straight
portion of a creep curve is shown in Fig. 3-12 reproducing the creep of a
0.35 per cent carbon steel at 850°F (454°C) after tests by F. H. Norton

1,000 1,500 2,000 2,500 3,ooo


TIME IN HOURS —»-
Fio. 3-12. The creep of 0.35 %C steel (K-20) at 850°F (454°C). Load 7,500 lb/in.'.
(.After tests by F. H. Norton.)

at a stress a = 7,500 lb/in.2 covering a testing time of 14,000 hr or nearly


2 years duration.
Finally, during the last period of creep the slope of the curve « = /(<)
starts to increase until fracture occurs. It is assumed that in the poly-
crystalline metals during the third creep period the grain boundaries
gradually break up because of the effects of long heating under continu
ous stressing, and the bar necks down.
26 THEORY OF FLOW AND FRACTURE OF SOLIDS

If a metal bar after it has crept is unloaded and the temperature main
tained, it contracts instantaneously by the strain e' = BC (Fig. 3-13).
The contraction may continue for a certain
time, however, as shown by the curve CD
in Fig. 3-13. This is the "recovery" which,
has already been mentioned. Presumably
the same phenomena occurring in the grain
structure during creep cause the "primary
"
creep as well as the "recovery." Unfor
tunately they have been given little atten
Fio. 3-13. Creep. tion in long-time creep tests in which almost
exclusively the minimum creep rates are
reported. Here then are types of permanent deformations which depend
on the time in which similarly the rates of strain are functions of the
time I even when the stresses do not change with time.
3-4. Elastic Hysteresis and Aftereffects. In this section we shall
finally mention some observations about the behavior of two groups of
materials. To the first belong materials in which are included small
inclusions of a soft or weak substance. The commonest example is cast
iron with its embedded particles of soft graphite. 1 Most porous materials
may behave similarly, e.g., concrete or sandstone. If a cast-iron cylinder
is tested in compression, one obtains within small values of the strains a
stress-strain curve OAB, as shown in Fig. 3-14. The shapes of such
curves and their applications in the theory of bending of cast bars have
been exhaustively investigated by C. Bach2 and others. Although the
first loading curve OAB in Fig. 3-14 reminds us of the curves we discussed
in Sec. 3-1, it is interesting to note that through a subsequent unloading
from the stress corresponding to the point A and reloading to the same
1 According to Bardenhf.uer (Der Graphit im Grauen Gusseisen, Stahl u. Eisen,
vol. 47, p. 857, 1947) gray cast iron contains from 2 to 3 per cent of graphite. If we
assume the specific weight of graphite as 2.1 and that of cast iron as 7.3, it will be
found that the 3 per cent by weight of graphite corresponds to 10.4 per cent by volume,
i.e., about one-tenth part by volume of cast iron is filled by the soft graphite particles
or veins. Under stress the soft graphite flakes or temper-carbon particles will mainly
deform within the cast iron. The harder constituents of the structure are displaced
like rigid bodies along the leaves of graphite. If these minute surfaces, along which
sliding occurs, are under pressure, the internal friction tends to prevent this sliding.
Under cyclic loading and unloading in cast iron as in materials with a similar structure,
considerable losses of energy must occur, and this is converted into heat. The areas
in the stress-strain curves under a loading and an unloading curve which represent the
stored energy and the one regained after a cycle is completed differ by a considerable
amount.
2
Bach, C, and R. Baumann, "Elastizitat und Festigkeit," 9th ed., Verlag Julius
Springer, Berlin, 1914.
ELASTIC AND PERMANENT DEFORMATION 27

stress a completely closed loop is obtained which has been shown in


Fig. 3-14 connecting the points A and C. The described cycle of loading
and unloading may be repeated an indefinite number of times with the
result that always the same loop is recorded.1 The curve OAB is called
the virgin stress-strain curve. If the test cylinder is reloaded from point
C to point A and the stress further increased, the curve continues along
the branch AB of the virgin curve.
Suppose, however, that after the specimen is loaded to point A and
completely unloaded to point C several cycles of loading and unloading
are carried out in each of which the
subsequent maximum or minimum
stress is changed in the manner indi
cated in the second figure of the loop
reproduced to the right of the first
figure. A spiral is thus obtained with
switchbacks of decreasing amplitudes.
The cusps in it are located along a
certain curve. If the amplitudes of
the loops are continuously decreased
in infinitesimal steps, the cusps exactly
reproduce the virgin curve of the
material. The curve shown in the Fio. 3-14. Elastic hysteresis.
second figure by a heavy line thus
coincides with the virgin curve after it is shifted in the new position so
that the origin 0 falls in the center of the spiral. This phenomenon is
called elastic hysteresis in analogy to the phenomenon of magnetic hystere
sis, which has striking resemblances to it and has been much studied by
physicists. We note that closed loops having smaller amplitudes arc
tilted with their axes and inclined more steeply the smaller their ampli
tudes are. The geometric properties of these loops in their relation to
the shape of the virgin curve constitute the phenomenon of elastic
hysteresis, which may be further studied after three additive portions
in the strains are defined:
a reversible strain, a deformation which must
depend on the succession of the applied stresses, and a third component
which may possibly depend on time.2
1 The
properties of these loops were first described in the dissertation of Berliner
(University of Gottingen, 1906). L. Prandtl, on whose suggestion this work was
carried out, subsequently analyzed them in several papers. The details in the text
above were reported by him in a seminar in a series of talks in Gottingen in 1921.
5
Closed loops, in contrast to the "drifting" of the loading and unloading curves,
which was mentioned for ductile metals, may be obtained for the latter if they are
tested under cyclic loading between a positive (tensile) and negative (compression)
stress of equal magnitude. Closed loops are obtained for an el astico- viscous sub-
28 THEORY OF FLOW AND FRACTURE OF SOLIDS

The second group of materials behaving in a peculiar manner may be


chosen from among the modern "plastics." The phenomenon to which
attention shall be called consists in the following. Suppose that a speci
men of a plastic is suddenly loaded by a force and this force is maintained
at a constant value for an indefinite time. At comparatively low tem
peratures, an elastic strain on loading will be observed, followed by
further strain or creep in the manner indicated in the left portion of
Fig. 3-15. After a sufficient time has elapsed, this additional "creep"
seems to cease completely. The con-
, verse effect is found after a sudden
unload
unloading, as shown in the right part
of Fig. 3-15. On unloading, the
specimen contracts instantaneously
by an elastic strain equal to the
E, .
ff
one it suffered on loading. Subse
f
quently it slowly contracts more until
this additional "recovery" finally completely stops. Further observa
tions have shown that these effects follow the law of superposition. Thus
a bar twisted first in one direction and shortly thereafter twisted in the
opposite direction may show afterstrains corresponding first to the type
of loading which was carried last, but after some time the aftereffects
of the first type of loading may also show up. The material thus appears
to possess a memory for the cycles of previous operations to which it was
subjected.
These elastic aftereffects were discovered by the physicist W. Weber
around 1835 and were thus known long before the other effects to which
reference has just been made. They were first observed in materials
which are known for their perfect elastic properties, as in glass at low
temperatures, in elastic springs (the tubes of Bourdon manometers), in
wires of fused silica in delicate instruments (galvanometers), in glass
thermometers, in silk filaments, etc.

stance in this manner. On the other hand, the fact that very narrow closed loops
may also be observed for loading and unloading cycles between a positive load and
zero load at normal temperatures for the ductile metals points to elastic hysteresis.
The "Bauschinger effect" may be related to these phenomena. Thin plates of a
ductile metal, if they are deformed permanently by bending under load, show a tre
mendous "hysteresis" in their loading and unloading curves after they are bent in the
opposite direction and bent back again. By such operations in the plate, plastic
strains and systems of large residual stresses are generated after each operation.
The latter observations must be attributed to nonuniform yielding in the microscopic
structures of solids ; the elastic and the plastic or viscous types of deformation must
occur simultaneously in them, but the latter are not uniformly distributed through the
grain structure.
ELASTIC AND PERMANENT DEFORMATION 29

In contrast to the creep of the metals at moderately high temperatures,


which continuously proceeds, this kind of slow creep seems to stop after
a certain time even though the load continues to act. The elastic after
effects produced by positive or negative changes of stress follow the law
of superposition, and the corresponding total strains add up with the
corresponding signs. Organic chemical compounds in which the mole
cules are arranged in long chainlike structures and in which the chemical
bonds are stronger lengthwise and weaker crosswise to the chains show
these effects in pronounced manner (plastics, rubber).1 These substances
have much smaller moduli of elasticity than the metals, although they
have surprisingly high strength (nylon). Thus, in contrast to the stronger
ductile metals, the permanent portions of their strains (the aftereffects)
are frequently of the same order of magnitude as the elastic strains.
This last property makes them act in recovery in a manner surprising to
observers who are more familiar with the plastic deformation of metallic
materials. The elastic aftereffect under cyclic deformation must also
cause looping.2
1 Concerning the mechanical behavior of similar materials reference is made to the
valuable monographs by It. Houwink, "Elasticity, Plasticity and Structure of Mat
ter," Cambridge University Press, London, 1937, 376 pp., and by H. Leaderman,
"Elastic and Creep Properties of Filamentous Materials and Other High Polymers,"
The Textile Foundation, Washington, D.C., 1943, 278 pp.
*
Comparative calculations of the energy loss per cycle and unit volume for given
types of these effects can be found in the valuable paper by G. H. Keuleqan, Statical
Hysteresis in the Flexure of Bars, Bur. Standards Tech. Paper 332, vol. 21, p. 145,
1925; also 365, vol. 22, 1938. Keulegan designates the hysteresis caused by the elastic
aftereffect pertinently as "hereditary hysteresis" in contrast to "static hysteresis,"
which is what we called "elastic hysteresis."

S"
CHAPTER 4

BEHAVIOR OF MATTER UNDER HIGH PRESSURE

It may be shown that the terms "solid" and "fluid" which serve to
differentiate the states of aggregation of matter are in many cases entirely
insufficient. This is especially true if we consider the behavior of mate
rials under high pressure. A knowledge of how matter behaves in the
solid or liquid states when exposed to high hydrostatic pressures is in
many respects of great interest. It can easily be shown that, in the
smaller and comparatively colder cosmic bodies like the earth, the pres
sure due to the weight of the outer layers increases with depth, up to
several million atmospheres at the center (1 atm = 1 kg/cm2 = 14.2
lb/in.2). In certain gaseous stars with enormous densities, of the type
of the dark companion of Sirius, pressures of thousands of millions and
possibly of thousands of billions of atmospheres probably exist.1 The
questions relating to the equilibrium in the colder cosmic bodies with an
aggregation of matter like that in the earth in which not only the pressure
but also the temperature increases considerably with depth, are of great
interest for both the astronomer and the geophysicist, as well as for the
geologist. In contrast with these enormous pressures it has so far been
possible to obtain maximum hydraulic pressures of the order of 30,000 atm
in the laboratory. However, it is noteworthy that even within this
range of pressure very remarkable changes in the properties of solids and
liquids may be observed, and these observations throw light on many
questions of interest to engineers and scientists.
The most reliable and comprehensive tests on which the present knowl
edge of the behavior of matter under high hydrostatic pressure is based
are those of Prof. P. W. Bridgman2 to which reference will be made in
the following.

1 According to James H. Jeans ("Astronomy and Cosmogony," p. 73, Cambridge


University Press, London, 1929; cf. also "The Universe around Us," by the same
author, p. 250, The Macmillan Company, New York, 1929), the faint companion to
Sirius is one of the most remarkable stars in the sky. The average density in this
comparatively small cosmic body has been estimated to be fifty to sixty thousand
times that of water. In this faint star, which belongs to a group of stars called the
"white dwarfs," matter is condensed to the utmost limit. Because of the high tem
peratures therein the atoms are thought to be almost completely broken up into their
constituent nuclei and electrons, both of which move about like the molecules of a gas.
2 The
author is much indebted to Prof. P. W. Bridgman of Harvard University,
Cambridge, Mass., for his kindness in furnishing complete reports of his extensive
experimental work on high pressure and related questions.
30
BEHAVIOR OF MATTER UNDER HIGH PRESSURE 31

4-1. Polymorphism. To illustrate what was said in the first two chap
ters we may say that many materials composed of one fundamental sub
stance only, such as chemical elements and compounds, have not one, but
several solid states. Among the elements exhibiting such modifications
of the solid state the following examples may be mentioned: sulphur,
below about 20,000 lb/in.2, is monoclinic, while above this pressure the
rhombic sulphur is in thermal equilibrium with the fluid.1 If we plot
the temperatures t at which the changes from one state to the other occur,
with varying pressures p, three equilibrium curves in the pressure-
temperature diagram will result, which (in the p, t plane, having p as
abscissas and t as ordinates) separate the three regions: monoclinic,
rhombic, and fluid sulphur. The three curves meet at a "triple point"
at 20,000 lb/in.2 and 152°C. Other well-known examples are iron (for
merly thought to have four modifications known as a, y, and iron;

/3,

S
these modifications, however, have probably but two kinds of crystals
having different structures or arrangements of the atoms in the space
lattice2), phosphorus,carbon (graphite, diamond).
In 1900, G. Tammann discovered two varieties of ice denser than the
usual kind. These he designated as ice and ice II. Since the funda
I

mental investigations of P. W. Bridgman,3 which covered wider range

a
of pressures, has been shown that water in the solid state occurs under
it

high pressure in five different modifications. In changing from one kind


of ice to another the specific volume changes abruptly. The kinds of ice
occurring under higher pressures are the denser. On the basis of careful
investigations relative to the behavior of water, may be mentioned
it

that the well-known lowering of the freezing point of water by pressure


observed only in the case of ice up to a pressure of 2,200 atm and a
is

temperature of — 22°C. At higher pressures than this the melting tem


perature of ice increases with the pressure. Bridgman has investigated
the melting curve (the bounding curve between the fluid and the solid
state in a diagram with the pressure as abscissas and with the temperature
as ordinates) of ice VI to above 20,000 atm and to temperature of
a

-r-76°C. The melting temperatures of ices III


to VI increase with the
pressure in similar way as do the melting temperatures of most materials
a

when the pressure increased. Ice VI melts at +26°C under 10,000 atm
is

pressure and at +76°C under 20,000 atm.


'Tammann, G.,
"
Aggregatzustande, die Zustandsanderungen der Materie in
Abhangigkeit von Druck und Temperatur," Leopold Voss, Leipzig, 1922.
'Wever, F., Stahleisen, vol. 45, p. 1208, 1925; Mitt. Eisenforschungsinst., Dussel-
dorf, vol. 1925; also Koerbeb, F., Schmelzen, Erstarren, Sublimieren, "Handbuch
7,

der Physik," Vol. 10, p. 192, Verlag Julius Springer, Berlin.


Bridgman, P. W., Thermodynamic Properties of Liquid Water up to 80°C and
'

12,000 kg/cm2, Proc. Am. Acad. Arts Sci., vol. 48, No. 1912.
9,
32 THEORY OF FLOW AND FRACTURE OF SOLIDS

4-2. Viscosity. The behavior of fluids with respect to viscosity under


high pressure is of considerable interest from a practical as well as a
scientific standpoint. As already mentioned, the viscosity of fluids mani
fests itself as a kind of inner resistance or internal friction, which opposes
each change of shape by the relative sliding of the small particles or layers
of fluid. If a solid sphere is allowed to fall through a liquid having con
siderable viscosity, after a certain time it will attain a constant speed.
The higher the viscosity of the liquid, the smaller will be the velocity
of the falling sphere through the viscous liquid under the force of gravity.
The coefficient of viscosity, as introduced by Eq. (2-2), page 11, can read
ily be determined by measuring the time of fall of bodies in liquids. It
is by this method that Bridgman1 has determined the viscosity of liquids
exposed to increasing pressures, relative to their viscosity at atmospheric
pressure. He found that the coefficient of viscosity increased rapidly
with the pressure. In general, the effect of pressure on viscosity is greater
than on any other physical property. The effect of pressure on viscosity
also varies greatly with the liquid. It must be expected that liquids
exposed to sufficiently high hydrostatic pressures will congeal so that they
become solids.
The relation between pressure and viscosity appears to be an exponen
tial function, i.e., if the pressure is increased by equal amounts the vis
cosity increases as the terms of a geometric series. The more complicated
the structure of the chemical molecule of the fluid, the more the viscosity
increases with pressure. For example, according to the investigations of
Bridgman, at 12,000 atm the viscosity for mercury increases by 30 per
cent over its value at zero pressure, and 2.7-fold for water, both at 30°C,
and from 10 times to 1,000,000 times its original value for very viscous
mineral oils. Kiesskalt2 recently has shown that at 1,000 atm the vis
cosity of mineral oils is six to twenty-two times as great as under atmos
pheric pressure. Under high pressures such as exist in the journals and
bearings of high-speed rotating machinery, e.g., steam turbines, the
change in the viscosity of the oil may become sufficient to affect its
lubricating qualities considerably.
As in other respects, water shows in this an exceptional behavior. At
low temperatures viscosity decreases at first with increasing pressure, but
this is an abnormal phenomenon and the curves soon tend again to rise
for greater pressures or higher temperatures.
4-3. The Compressibility of Metals. The volume of solids and fluids
decreases slightly under pressure. This decrease in volume is usually,
1 Proc. Am. Acad. Arts Set., vol. 61, p. 86, 1925; vol. 62, p. 187, 1927; compare also
tests by Faust and Tammann, referred to in the book by the latter, mentioned above.
* Mitt. u. Forschungsarb. Ver. dent. Ing., vol. 291, 1927.
BEHAVIOR OF MATTER UNDER HIGH PRESSURE 33

at small pressures, so small that it may be detected only by the aid of


sensitive measuring instruments. It is also a perfectly reversible or an
elastic change, since, if the pressure is removed, most solid and liquid
materials take on their initial volume. An exception to this is the case
of porous materials (e.g., wood, cast iron) which can be permanently com
pressed by high hydrostatic pressure.
In order to give an idea of the compressibility of metals a few numerical
data will now be given. The investigations of Bridgman1 show that
for iron the ratio of the decrease in volume At; to the initial volume v<>(the
initial volume taken at 1 atm pressure and some standard temperature,
which was 30°C in Bridgman's tests) is equal to:
For iron
— = 10-7(5.87 - 2.10 X 10-*p)p ,

where p is the pressure in kilograms per square centimeter. A more


exact definition of the compressibility k is the ratio of the decrease in
volume produced by a pressure increase of 1 atm or dv/dp to the volume v
existing at the time considered:
,
K — - -r-
1 dv
.
V dp

For most materials the compressibility k decreases gradually with


pressure; in other words, compressed metals, for the same increase in
pressure, change in volume a smaller amount, the higher the compression.
The measurements of Bridgman give (pressures p are in kilograms per
square centimeter) :
For copper

00
= 10-7p(7.32 - 2.7 X 10-'p) .

For aluminum
Av
I'o
= 10-7p(13.34 - 3.5 X 10"»p) .

For lead
Av
»1
= 10-7p(23.73 - 17.25 X 10-»p) .

The greatest compressibilities are exhibited by the alkali metals. Of


all elements, cesium is the most compressible, the volume of cesium
decreasing by one-third at 15,000 atm pressure.
As one may see from the foregoing the volume of steel or iron decreases
1 The
Compressibility of Thirty Metals, Proc. Am. Acad. Aria Sci., vol. 58, No. 5,
p. 166, 1923.
3-1 THEORY OF FLOW AND FRACTURE OF SOLIDS

only about 0.0006 of its initial volume at a pressure of 1,000 atm. Conse
quently, as a rule we may neglect such changes in volume due to pressure,
when considering the deformation due to plastic flow which may occur in
ductile metals.
On the other hand, compressibility plays an important part if the state
of equilibrium of the rocks at great depths in the earth is investigated.
Because of the effect of pressure the density of these rocks must be con
siderably increased when compared with their density at the surface of
the earth. In connection with possible applications of these facts to
geophysical questions some results of tests of Bridgman on nonmetallic
materials will be mentioned here.1
4-4. The Compressibility of Artificial and Natural Glasses.1

Silica a = 27.74 X 10"T b = 7.17 X 10-"


Pyrex glass a = 30.08 X 10"' 6 = 4.86 X 10-"
Tachylite Kilauea (Hawaii) a = 13.74 X 10-7 6 = -9.1 X 10"1!
The compressibility is given here at 75°C as Av/v<> = ap + op*; the constants a and 6
are given for p in kilograms per square centimeter. It is remarkable that with the
exception of the last example the coefficient 6 is positive and hence these glassy sub
stances are the more compressible the higher the pressure. This abnormal behavior,
found only in glassy substances, seems
apparently to be connected with the silica
content.
4-6. The compressibility of liquids is
in general much greater than that of solids
and amounts at 12,000 atm to about 20
per cent for water and about 30 per cent
for ether, one of the most compressible
4 6 8 10 epOOATM, liquids. For most liquids the compress
PRESSURE p — «- ibility decreases much more rapidly at
Fia. 4-1. Compressibility of ether. 'Ac the lower pressures than at the higher.
cording to P. W. Bridgman.) A curve typical of most liquids is shown
in Fig. 4-1. This curve is for ether and
is reproduced from the tests of Bridgman.
4-6. The Effect of Pressure on the Rigidity of Metals. By measuring the relative
movement of two helical springs, one consisting of steel and the other of the metal to
be tested, which were immersed in a liquid in a pressure vessel, P. W. Bridgman* was
able to determine the relation of the moduli of rigidity. The two springs were
attached to each other, and the effect of pressure on the modulus of rigidity of the
steel spring had previously been determined by absolute measurements. The
1 The
compressibility of minerals and rocks with special reference to problems of
geophysics has been investigated particularly by L. H. Adams and his collaborators in
the Geophysical Laboratory at Washington, D.C. These tests will be referred to in
Vol. 2, where further experimental data on the elastic compressibility of rocks will
be given.
•Bridgman, P. W., Am. J. Sri., vol. 10, p. 350, 1925.
'Proc. Am. Acad. Arts Set., vol. 64, No. 3, p. 39, 1929.
BEHAVIOR OF MATTER UNDER HIGH PRESSURE 35

modulus of rigidity under a pressure of 10,000 kg/cm' was increased above its value
at zero pressure by the following amounts:

%
Tantalum . . +0.3
Platinum . . . +2.4
Nickel + 1.8
Spring steel +2.2
These results may be of importance, if the influence of the tremendous
pressures in the central core of the earth on the velocity of propagation
of the earthquake waves passing through it is to be considered. The
velocity of the distortional waves depends on the modulus of rigidity
and, since the pressures in the central part of the earth reach millions of
atmospheres or so, the effect of pressure on rigidity at these great depths
may therefore be appreciable.

For further information the reader is referred to P. W. Bridoman; "The Physics


of High Pressure" (international textbooks of exact science), The Macmillan Company,
New York, 1931, 398 pp.

4-7. Cohesion in Liquids. A few remarks at this point may be in


order, although their relation to the behavior of matter under high pres
sure is not immediately obvious. A. A. Griffith1 has shown that very thin
filaments of silica glass were much stronger above their softening temper
ature than in the cold condition. At first glance this is the opposite of
what would usually be expected. From capillarity phenomena it is
known, however, that very thin films of liquids may sustain extremely
high tensile stresses. From these and numerous other considerations it
must be concluded that in fluids under suitable conditions a considerable
resistance against tearing apart (the effect of the cohesive forces acting
between the atoms or of the "cohesion") is observed, a property which
usually is ascribed only to materials in the solid state.
4-8. Optical and Other Effects. By means of a skillfully designed
arrangement which consisted of a special steel pressure cylinder of an
inverted T shape containing a movable piston, two glass or quartz win
dows, and a simplified packing Thomas C. Poulter2 succeeded in observ
ing various effects, due to high hydrostatic pressures up to 30,000 atm, on
material properties, e.g., upon the index of refraction (of glass and of
paraffin), upon the rotation of the plane of polarization of optically
active compounds in solution (sugar), and upon the destruction of bac-
1 Trans. Roy. Soc. London, set. A, vol. 221, p. 163, 1921.
2
Apparatus for Optical Studies at High Pressure, Phys. Rev., vol. 40, pp. 860-876,
1932. J.
See also Applied Phys., vol. 9, p. 307, 1938. Many investigations on the
effect of high hydrostatic pressures have been carried out under T. C. Poulter at the
Armour Research Foundation in Chicago, 111.
36 THEORY OF FLOW AND FRACTURE OF SOLIDS

teria. He and L. Uffelman1 investigated the penetration of hydrogen


through the walls of a very fine grained steel cylinder and found a rapid
evolution of hydrogen at pressures as low as 4,000 atm through a thickness
of steel wall of % in. after an exposure of 5 min to pressure. Alcohol,
ether, and water were shown2 to penetrate small cylinders of glass or
fused quartz to considerable depths if the pressure around them was
raised beyond any value above 15,000 atm and maintained for as long as
15 min, after which the pieces shattered if the pressure was rapidly
released. However, if the pressure was quite slowly released over an
extended period of several days, no breakages occurred. Oil and glycerin
did not cause fracturing under the same conditions. Some further obser
vations on the penetration of a liquid into the fine fissures which are
present in brittle materials, such as porcelain or natural rocks, when a
high hydrostatic pressure is exerted on such specimens in lateral direction
through the liquid will be mentioned in Chap. 15, page 206, and Chap.
17, page 244, causing premature fractures in such tests.
4-9. Munroe Effect. Some very interesting observations on extremely
highly concentrated pressures exerted when a metal jet impinges on a
SHOCK WAVE
TARGET

LINER
DETONATOR

STEEL LINER'
Fig. 4-2. Explosive charge. Fig. 4-3. Formation of a needle
of steel from conical liner.

metal block were described3 in connection with what became known as the
Munroe effect of a high explosive charge. If a cylindrical charge of an
explosive substance placed against a solid block of steel is hollowed out on
the side in contact with the block according to a cone and is covered by a
thin steel sheet (metal liner) in the form of a conical shell on the side with
the cavity in the charge, some very strange effects can be observed when
the charge is fired by means of a detonator placed on the opposite end of
the cylindrical charge (see Figs. 4-2 to 4-6). When the charge is exploded
a plane shock wave moves to the right through the charge. After this
wave has passed the vertex of the steel cone the portion of the metal liner
1
Physics, vol. 3, p. 149, 1932.
1 Poulter, T. C, and It. O. Wilson, The Permeability of Glass and Fused Quartz
to Ether, Alcohol and Water at High Pressures, Phys. Rev., vol. 40, p. 877, 1932.
3 Birkhoff,
G., D. P. MacDougall, Emerson M. Pitch, and Sir Geoffrey
Taylor, Explosives with Lined Cavities, J.
Applied Phys., vol. 19, No. 6, pp. 563-582,
1948.
BEHAVIOR OF MATTER UNDER HIGH PRESSURE 37

beyond which the shock wave has advanced is moved inward with
extremely high velocities. Because of the radial velocity field the dis
placed masses of the metal liner assemble around the axis of the cone
and form a thin rod (needle) of steel which is shot with a very high veloc
ity into the solid metal block.1 The effect is
similar to that if a fine jet of fluid steel having
the velocity v were directed toward the metal
block penetrating the target with a relative

V&W//A
Fio. 4-6. Fia. 4-6.
Fias. 4-6 and 4-6. Solid
Fio. 4-4. Steel jet penetrating cylinder of steel (left) and
metal block. of lead (right) after being
pierced by steel jet of metal
liner of charge shown in Fig.
4-2.

velocity u. Thus by Bernoulli's theorem, assuming that the material in


the needle behaves like a fluid in stationary flow relative to the contact
area between the steel jet and the metal block, the pressure in the center
of the contact area is equal to

~
P = {V U)J
2g

where y is the specific weight of the material in the jet and g the gravity
acceleration. The authors reported having reached "cutting pressures"2
of 280,000 atm. The effect of the penetrating steel jet is reproduced in
Figs. 4-5 and 4-6 for a block of steel and a block of lead.
1 This action of a jet of steel was used in the antitank projectiles of the bazooka for
piercing armor plate. The head of such a projectile has an "ogive," a rounded por
tion ahead of the conical cavity in the charge to reduce its air resistance.
* It is known that steel plates can be cut
through the action of a high-velocity
water jet at room temperature.
CHAPTER 5

THE ORDERED AND UNORDERED STATES OF MATTER


So far we have classified materials according to their state of aggrega
tion, i.e., "solid" or "fluid," depending on their behavior under various
states of stress. We now consider a different mode of classification,
namely, that given by their internal structure. According to G. Tam-
mann,1 the anisotropic state of matter should be distinguished from the
isotropic. The elementary particles of matter (atoms, ions, molecules)
out of which all materials, according to the modern view of physics, are
composed, are, in the first (the anisotropic) state arranged in a definite

Fig. 5-1. Fio. 5-2.


Fios. 5-1 md 5-2. Slip hands produced by impression of a needle on a copper crystallite.
Fig. 1. Impression on a cubic plane. Fig. 2. On an octahedral plane. (A. Mailer.)

geometrical and regular way; while in the second (the isotropic) state,
they are unordered. To the first state belong the crystals and the crystal
line materials; to the second state belong the gases, liquids, and vitreous
(solid or fluid) materials.
The view of an ordered arrangement of the elements of crystals is sup
ported by many and varied observations, so that we may take it as almost
a certainty. These views were originally suggested from the geometri
cally regular shape of natural crystals, those precipitated from solutions,
or from the molten state. The best proof of the correctness of these views
however, given by the interference observations of X rays2 passing
is,

" Aggregatzustiinde, etc.," Leopold Voss, Leipzig, 1922.


1

Ewald, P. P., Der Aufbau der festen Materie und seine Erforschung dureh
*

Rontgenstrahlen, "Handbuch der Physik," Vol. 24, Chap. p. 191; Born, M., and
4,

O. F. Bollnow, Theoretische Grundlagen, ibid., Chap. p. 370; Bragg, W. H., and


5,

W. L. Bragg, "X-rays and Crystal Structure," George Bell Sons, Ltd., London,
&

1924 Wyckoff, R., "The Structure of Crystals," Chemical Catalogue Company, Inc.,
;

New York, 1924; Gi,ocker, R., " Materialpriifung mit Rontgenstrahlen," Verlag
Julius Springer, Berlin, 1927.
38
THE ORDERED AND UNORDERED STATES OF MATTER 39

through crystals. Other observations in support of this view are the


extraordinarily regular markings which one may produce upon the pol
ished surface of crystals by etching (etching figures), by slow plastic

o
strains (slip lines), or by impact strains (impact figures).1

Fig. 5-3. Fig. 5-4. Fig. 6-5. Fig. 5-6.


Fios. 5-3 to 5-6. Etching figures on spheres cut from copper single crystals. These
figures were obtained by using different solutions and are reproduced schematically. On
the surface of the etched copper balls a figure appears, which shows the contour lines of a
cube in Fig. 5-3, of an octahedron in Fig. 5-4, of a rhombic dodecahedron in Fig. 5-5, and
the combination of a cube with an octahedron in Fig. 5-6. (According to K. W. Hausser
and P. Scholz.)

In vitreous solids, on the contrary, no orderly arrangement of the


atomic elements can be detected by X-ray methods. This is also the
1
Anton MOller in the Tammann Institut (" Metallographische Studien," Dis
sertation, Gottingen, 1926) has shown that extraordinarily regular figures may be
produced by pressing a needle on the crystals of metals.
The space lattice of copper, aluminum, silver, gold, and y iron is a face-centered
cubic, that of or, and iron
/3,

body-centered cubic, while that of zinc hexagonal


is

is
a
5

a
lattice. The elementary cubes of the former contain an atom in each corner and the
middle of each side. In the body-centered cubic lattice there also an atom in the
is

middle of the cube, besides that in the corners. In zinc crystals, the atoms have
approximately the same arrangement as the middle points of balls when piled in a
large heap in the densest manner possible.
In a copper crystal certain definite crystallographic planes are the planes of a
cube, an octahedron, and rhombic dodecahedron. The octahedron results we
if
a

join the atoms of the middle of each side of a cube while the planes of the rhombic
dodecahedron lie oblique to the 12 edges of the cube.
According to Mttller, a needle pressed upon a crystal of copper there pro
if

is

is

duced in the neighborhood of the impression the system of lines shown schematically
in Figs. 5-1 and 5-2. If the needle pressed upon a plane of the cube the slip lines
is

of Fig. 5-1 result while pressed upon an octahedral plane the lines of Fig. 5-2 are
if

produced. Under plastic deformation the octahedral planes of the copper crystals
slide the most easily; hence their traces on the polished surface appear under the
microscope as the system of lines of Figs. 5-1 and 5-2. Other observations on figures
produced by pressure and etching are given in "Moderne Metallkunde" by Czochralski.
On large balls, about the size of a billiard ball, cut out of single crystals of copper,
Hausser and Scholz (Wiss. Veroffentl. Siemens-Konzern, vol. Part 1927) have
3,
5,

made very remarkable observations. They obtained, by an etching process, the sys
tem of lines on the surface of the balls shown schematically in Figs. 5-3 to 5-6. The
lines in these figures indicate curves where the mode of reflection changes on the
polished and etched surface of the balls. Hausser and Scholz were able to produce,
on the surface of balls consisting of single crystals of copper (or silver), a system of
lines with a cubic, a pure octahedron, or a pure rhombic dodecahedron symmetry,
depending on the particular etching agent or etching procedure.
40 THEORY OF FLOW AND FRACTURE OF SOLIDS

case with fluids.1 With the exception of monomolecular films, we may


therefore regard vitreous solids as undercooled fluids.2 Whether or not
the elementary particles of liquids and vitreous substances are completely
can by no means be stated easily. What we know that

is,
unordered
ordinary liquids certainly are not united in an orderly arrangement as
are the elements of crystal lattice. In fluids or vitreous substances

a
these elementary particles may, however, form long chains or layers
fastened together in such a way that at higher temperatures they may,
more or less easily, bend or slide over one another in all directions. On
the basis of certain observations, A. A. Griffith' was led to this view.4
Such theory also supported on the basis of capillarity phenomena by
is
a

the enormous value of cohesion forces, observed in very thin films of


liquids, in very viscous quartz filaments, and in the surface layers of
fluids. For example, Griffith took hardened steel ball, such as used

is
a
in ball bearings, and fitted to very carefully made hole in a plate, such
it

a
that there was only an extremely small amount of clearance between
the ball and the plate. The ball would easily fall through the hole,

it
if
were dry; however, when drop of water was placed in the annular space
a

between the ball and plate, the ball stuck in the hole and could be forced
through only by applying large force. According to Griffith, the mole
a

cules of water fasten themselves in chains to the surface of the ball and
the hole. These chains reach across the circular space between the ball
and the hole and by virtue of their great tensile strength hold the ball
fast.
Liquid helium which boils at — 269°C (4°C above absolute zero)
I

under atmospheric pressure when further cooled undergoes a transition to


helium II, at 2.2° abs. Helium II has been described5 as a fluid without
viscosity which creeps through vacuumtight joints and up along the walls
of container in an invisible film in the direction opposite to gravity.
a

Its thermal conductivity greater than that for any other substance.
is

He II conducts heat as wave motion in the same way that fluid trans-
a
a

In "Die Lehre von den fliissigen Kristallcn" (Bergmann, Wiesbaden, 1918)


1

Ldhmann describes "flowing" and "fluid crystals," using certain substances having
a

complicated molecular structure but which we may regard as fluid. However,


although they may be shown optically to be anisotropic like crystals, X-ray analysis
showed that they have no lattice structure (see E. HOckel, Dissertation, Gottingen,
1921).
Tammann, G., "Aggregatzustande, etc." Leopold Voss, Leipzig, 1922.
2

Trans. Roy. Soc., London, scr. A, vol. 221, p. 163, 1921.


'

spontaneous parallel arrangement of the molecules of a fluid explained by the


A
4

is

electric or magnetic moments produced by these molecules (see E. HOckel, Disserta


tion, 1921; and Born: Ber. Berl. Akad., 1916).
But. Standards, Tech. News Bull., vol. 33, No. p. 13, February, 1949.
2,
5
THE ORDERED AND UNORDERED STATES OF MATTER 41

mits elastic dilatation and compression waves as sound. The remarkable


properties of He II perhaps disclose a new ("fourth"?) state of matter.
It was mentioned that crystalline substances in general increase their
specific volume when they pass from the solid to the fluid state at the
melting point. In the gaseous state the atoms or molecules are far apart
and assumed moving rapidly in space; in the fluid state their distances
must, however, be comparable with the spacing they possess in their
regular arrangement in the crystal lattices. To account for the com
paratively small difference of the specific volume of a substance in its
crystalline and in its molten state and for a statistical treatment of the
phenomena of diffusion and viscosity in liquids H. Eyring1 assumed that
just as a gas consists of atoms or molecules moving about in empty space,
a liquid may be regarded as made up of "holes" or empty spaces moving
about at random in "matter." Conditions may be roughly compared
with those of the sand grains deposited through the wind in a dune and
in consolidated condition on a beach. In the former, they are "loosely,"
in the latter "densely" packed, corresponding to the liquid mobile and
to the solid regularly packed states of the atoms. The atomic particles
in a liquid are surrounded by nearly as many neighbors as in a solid, but
in many small regions they may still be partly bound in regularly arranged
groups. A glass in the solid state not showing signs of a regular arrange
ment of the molecules (in X-ray diffraction patterns) after some reheating
below a temperature at which it would soften considerably may slowly
"de vitrify" into small groups (shown by the X-ray analysis) having signs
of some regular arrangements, but which cannot yet be classified as
developed crystallites. The thermal agitation of the molecules in the
fluid state facilitates the jumping away of the molecules from these bound
groups into the "holes," filling the old ones and creating new ones.
Observations by X rays on the fine structure of the organic amorphous
materials (high polymers, nylon, plastics, rubber) also indicate that in
these substances made up of long chains of molecules some orderly
arrangement in small regions exists, in the "micelles" showing a quasi-
crystalline structure. It is assumed that the long chains have their
strongest chemical bonds in their longitudinal, and weaker bonds in the
lateral directions. Through the latter the chains over portions of their
length may be joined closely and form regular bundles with a quasi-
crystalline structure, while the open ends of the chains are irregularly
twisted around each other. Under a high tensile stress more of these
loose ends may be brought in a parallel direction and nearer to each
1 Eyrino, Henry, Samuel Glasstone, and Keith J. Laidler, "The Theory of
Rate Processes," Chap. 9, p. 477, McGraw-Hill Book Company, Inc., New York,
1941, 611 pp. Also Eyrino, H., /. Chem. Phys., vol. 4, p. 283, 1936.
42 THEORY OF FLOW AND FRACTURE OF SOLIDS

other, making the structure in the highly strained condition still more
"oriented."1
1 Leaderman, Herbert, "Elastic and Creep Properties of Filamentous Materials
and Other High Polymers," p. 91. The Textile Foundation, Washington, D.C,
1944, 278 pp. The connection between the molecular structure and the mechanical
properties of the so-called high polymers is a preferred subject of modern physico-
chemistry. Cf. the illustrating talk by Herman Mark (Recent Trends in Polymer
Chemistry, Chem. Eng. News, vol. 27, pp. 138-143, 1949), in which mechanisms of the
chain models of cellulose, isoprene, and other similar organic substances are described.
CHAPTER 6

CRYSTALLINE STRUCTURE IN METALS


In the metals commonly used as materials of construction the smallest
parts having uniform properties, namely, the crystal grains, are at most
of very small dimensions compared with the dimensions of the material
bodies. The mean diameter of these grains is at most of the order of a
few millimeters; usually it is only to M"oo mm. In contrast to this
the distance between the atomic particles of a crystal lattice is of
the order of 10-8 cm. The study of the fine crystalline structure of the
metals and their alloys with the aid of the optical and the electron micro
scope has disclosed important information concerning its influence on the
strength properties of the metals and also relative to the visible changes
which can be observed in the grain structure accompanying the
phenomena of the plastic deformation of solid metals or causing their
fractures. As a rule a metal with a very fine grain has in general a
greater strength than the same metal with a coarse granular structure.
Since the grain size and the conditions in the crystalline structure of a
metal can be greatly changed by suitable handling during its manufacture
and by mechanical and heat treatment, it is obvious that these metal
lurgical factors must be of the greatest importance in their influence upon
the properties determining the mechanical strength of metals. As far
as these factors do not lend themselves to an analysis by means of the
principles of mechanics, however, they cannot be considered here, and
reference regarding them must be made to treatises on physical metal
lurgy.1 They will be reviewed very briefly in the following.
1 Tammann, G., "The States of Aggregation" (translated by R. F. Mehl), D. Van
Nostrand Company, Inc., New York, 1925. Seitz, Frederick, "The Physics of
Metals" (Metallurgical Engineering Series), McGraw-Hill Book Company, Inc.,
New York, 1943, 330 pp.; "The Modern Theory of Solids" (International Series in
Physics), McGraw-Hill Book Company, Inc., New York, 1940, 698 pp. Boas, W.,
"An Introduction to the Physics of Metals and Alloys," Melbourne University Press,
Victoria, Australia; also John Wiley & Sons, Inc., New York, 1947, 193 pp. Ober-
hoffer, P., "Das technische Eisen," 2d cd., 1925; Czochralski, J., "Moderne
Metallkunde in Theorie und Praxis," 1924; Koerber, F., Schmelzen, Erstarren,
Sublimieren, "Handbuch der Physik," Vol. 10, p. 192, Verlag Julius Springer, Berlin.
Sachs, G., "Praktische Metallkunde," Verlag Julius Springer, Berlin, Vol. 1, 1933,
272 pp.; Vol. 2, 1934, 238 pp.
43
44 THEORY OF FLOW AND FRACTURE OF SOLIDS

In engineering constructions the pure metallic (chemical) elements are


comparatively little used with the exception, for example, of electrolytic
oxygen-free copper in electric machines or aluminum in conductors, etc.
Through very minute amounts of additional constituents the strength
properties of metals can be considerably changed; for example, in ordi
nary low carbon steel a few hundredths per cent carbon raises the low
yield point of a pure iron to 30,000 lb /in. J. Metallic specimens either are
composed of one kind of crystallite, as in the pure metallic elements or
in metals which contain very small amounts of foreign constituents, or
two or more metallic elements are combined in alloys in comparable
proportions which may or may not contain crystal grains of one lattice
structure. The principles of the formation of the crystallites from the
liquid state in binary (two-component) alloys have been exhaustively
analyzed by means of the phase diagrams in which the temperatures are
plotted as a function of the fractional ratios of the components. The
magnitude of the forces under which a metal will yield plastically or frac
ture depends on its composition and crystal structure, on whether it
consists of small or large grains of one or more kinds, and on the strength
of the intercrystalline grain boundaries and their constituents. The size
of the crystal grains in cast metals and their mode of formation depend on
the rate of cooling during the solidification. The absolute size of the
grains, on the other hand, may be changed in solid metals by a cold
deformation followed by a subsequent reheating. This is known as
recrystallization. The amount of previous cold deformation and the
temperature during the reheating and its duration determine the absolute
size of the grains after recrystallization. As a rule recrystallized grains
are larger the smaller the permanent strain was to which the metal was
prestretched cold and the higher the temperature during the subsequent
heating was, assuming otherwise equal conditions (equal times of heat
ing).1 After the First World War, this made possible the artificial grow
ing of large metallic single crystals in sizes of specimens used in tensile

1 Probably one of the first systematic investigations on grain growth in solid metals
was made around 1916, by J. Czochralski, who proposed a "recrystallization dia
"
gram in which the strain of prestretching and the temperature during the subsequent
heating were plotted as rectangular coordinates and the average grain size (mean
diameter of the recrystallized grains) as ordinates. (His paper in Z. Metallkunde,
vol. 19, p. 316, 1927, contains a historic review of this work.) He produced the first
large single crystals of aluminum. Concerning recrystallization in metals see G.
Sachs, "Mechanische Technologie dcr Metalle," p. 170, Leipzig, 1925, 319 pp., and
E. Schmid and W. Boas, "Kristallplastizitaet," p. 217, Verlag Julius Springer, Berlin,
1935, 373 pp.
CRYSTALLINE STRUCTURE IN METALS 45

testing machines, which have offered excellent new means for a better
determination of the mechanisms of plastic distortion in metals.1
Small amounts of nonmetallic or metallic elements in a pure metal can be present
in finely dispersed condition throughout the lattice of the crystal grains, or the former
may form chemical compounds with the metal, which may occur in the shape of hard
or soft minute inclusions, globules, needles, etc., or these compounds may be distrib
uted along the grain boundaries. One constituent of steel is perlite, which consists of
fine intermittent layers of iron carbide known as hard cementite and of soft a iron.
The perlite in a low-carbon steel may assemble in microscopic chunks within or near
the ferrite In normalized steel with equalized small ferrite grains
(or

iron) grains.

a
it probably favors a distribution along the grain boundaries.
The shapes and sizes of microscopic particles of foreign substances in a metal can
frequently be controlled through the proper "heat treatment"; a parallel orientation
of nonmetallic inclusions in a structure may be due to rolling or drawing. In low-
carbon steel a finely or coarsely lamellated perlite might be obtained by application
a

of heat and the proper cooling rates. Pure a iron (the ferrite crystals in low-carbon
steel) very soft readily deformable metal. It probable that the high value of

is
is
a

well and sharply denned yield point which observed in the tensile tests of a normal
is

ized (annealed slightly above the critical temperature of 906°C at which the a iron
ceases to exist) mild steel consisting of a iron with a few hundredths of per cent

1
carbon and small traces of manganese, silicon, etc., caused by the minute hard
is

particles of the cementite which are assembled along the boundaries separating the
small ferrite grains. Since the otherwise easily deformable ferrite crystals are sur
rounded by a hard skeleton, they will not deform under smaller stresses. The strong
boundaries inhibit their deformation until the stresses reach high values which are
large enough to break up the resistance of the grain boundaries. This then observed

is
as a discontinuity in the stress-strain relation and as a yield point of the composite
structure.* In other cases, however, small quantities of a weak foreign substance,
assembled in the grain boundaries, may have the opposite and an injurious effect, as,
for example, with iron sulfide in steel3 or with copper oxide in copper, and in the latter
cases the strength decreased through the boundary substance. The grain bounda
is

ries in a polycrystalline metal in general must exert restricting effect on the deform-
a

ability of the crystallites even in pure metals. This has been demonstrated in tension
tests on specimens composed of just a few large crystallites of aluminum. When such
specimens are pulled, they do not stretch uniformly but disclose knots and bulged
portions at larger strains along the few grain boundaries that separate the large grains.
In these cases the boundaries do not contain a foreign substance, and the misalign
ment of the crystallographic planes of slip in neighbor crystals causes a great disturb
ance in the strains.

Another single metal crystals consists in permitting the


method for producing
1

growth of tube connected with the


one crystal from a nucleus situated in a capillary
molten metal and in slowly moving the glass tube containing the liquid metal into
cooler positions, maintaining the required temperature gradient for crystallization at
the advancing front of the slowly growing crystal.
Schweiz. Bam., vol. 83, Nos. 14, 15, 1924. See also Chap. 18, p. 285.
«

Goerens, P., "Ueber Stahlqualitaeten und ihre Beziehungen zu den Herstellver-


*

fahren," Krupp. Monatsh., 1927.


46 THEORY OF FLOW AND FRACTURE OF SOLIDS

An instructive example of the changes in the structure of a medium-


carbon steel (0.16%C) was described by R. W. Bailey1 in his early pioneer
work on the creep of this metal at high temperatures. From a long
superheater tube of a steam boiler that had been exposed to high temper
atures during 16,000 hr micrographs taken in a number of positions along
the tube were investigated by Bailey. Owing to the construction, these
positions had been exposed to progressively increasing temperatures
along the tube. Two micrographs shown in Figs. 6-1 and 6-2 are
reproduced from his paper. Figure 6-1 shows approximately the initial

Flo. 6-1. Structure of a 0.16%C steel. Fio. 6-2. After splieroidization of the per-
The dark areas are perlite. lite. (Magnification 235.) (After R. W .
Bailey.)

structure in an annealed medium-carbon steel; the dark grains are


perlite. Figure 6-2 shows the structure of the steel after long exposure
to steam temperature (the highest was 340°C). The lamellar perlite
grains have entirely disappeared, and the cementite in the perlite has been
broken up into many small globules. (The edges of the polyhedral
crystallites in metals may be rounded when the temperature is sufficiently
high to soften their interior so that the surface tension may deform
them like the droplets of a fluid.) The structure shown in Fig. 6-2 is
well known to metallurgists as one indicative of "spheroidized
"
perlite.
The time required to spheroidize the cementite is known to depend
on the temperature. Incidentally, Bailey determined this time experi
mentally and found it to follow the function t = cei/T where t denotes the
time, T
the absolute temperature, and b, c are constants.
Metals and their alloys cast from the molten state upon their solidifica
tion frequently show an oriented structure. In ingots the crystallites
grow faster in the direction parallel to the largest gradients in the tem
perature field during the cooling of the metal than in other directions.
Crystallites forming long needles with many branches are known as
1 "Creep of Steel under Simple and Compound Stresses and the Use of High Initial

Temperature in Steam Power Plant," paper presented at the World Power Conference,
Tokyo, 1932.
CRYSTALLINE STRUCTURE IN METALS 47

dendrites. Two examples of dendritic oriented grain structure in an


a
alloy are illustrated in the micrographs reproduced in Figs. 6-3 and 6-4.

Fig. 6-3. Dendritic structure of "refraet- Fio. 6-4. Dendritic structure of "rcfract-
alloy" (Westinghouse) in cast condition alloy" (Westinghouse) in cast condition
serving for high-temperature applications. serving for high-temperature applications.
(Magnification 50.) (Acknowledgments are (Magnification 50.) (Acknowledgments are
due to H. ScoU and E. F. Losco.) due to H. ScoU and E. F. Losco.)

Fio. 6-5. Microstructure of cast "refractalloy " after 44 hr aging at 1700°F. (Magnifi
cation 50.) (H. ScoU and E. F. Losco.)

The minute quantities of gases (hydrogen, nitrogen, oxygen) which may he present
in solid solution in metals can affect their strength properties. Commercially pure
copper, which contains traces of dissolved oxygen, if annealed in a hydrogen atmos
phere (for preventing scaling) becomes entirely brittle. The atoms of one metal
may diffuse through the lattice of the crystals of another metal. One practical use
of this is the joining of steel parts by copper brazing. Copper in the molten condition
is drawn into the small clearance between two steel parts by the capillary action. It
has been found, for example,1 that two steel bars can be joined together by a layer of
very pure oxygen-free copper of 1/1,000 in. (or less) thickness so solidly that a plane
joint will have a strength of 120,000 to 130,000 lb /in.' in tension perpendicularly to
the plane at separation. (Soft annealed pure copper itself has a tensile strength of
33,000 lb /in.*.) The high strength of steel is imposed on the thin copper layer not by
mechanical reasons (the two adjoining steel parts of higher strength prevent the lateral

1 Unpublished report by M. J. Manjoine.


48 THEORY OF FLOW AND FRACTURE OF SOLIDS

plastic contraction in the thin copper layer stressed in tension) but by diffusion of the
copper atoms into the lattice of the crystallites of the iron (creating atomic adhesion)
and of those of the iron into the lattice of the copper grains (reinforcing the weaker
copper against plastic slip).
Alloys of two miscible metals may have a maximum strength at some definite pro
portion of the two components, and the strength may be much higher for this combi
nation than for each of the pure components. The optimal strength may sometimes
occur after a very small amount of a second metal is introduced to a pure metallic
clement, e.g., the addition of just a few ounces of silver to a ton of oxygen-free pure
copper improves the long-time creep strength of copper at a temperature of 120 to
150°C (lowers the small rate with which copper stretches continuously under stress
to a minimum value at these temperatures). Optimal strength and greatest hardness
values in alloys are achieved by exposure to heat followed by the required rates of cool
ing, including a very rapid one (quenching). There are two further important
purposes for the application of heat: annealing for stress relieving (usually at moder
ately high temperatures) and recrystallization combined with a previous cold deforma
tion. Through the former the undesirable or deleterious systems of inherent or
residual stresses are relieved (this is an application of the process of the relaxation
which was mentioned in Chap. 1, page 4) which are caused by the various processes
applied in the manufacture of the raw metallic materials or during their subsequent
machining operations. Residual stresses are caused by thermal stresses through
unequal heating or cooling (in cast or welded pieces) or nonuniform permanent dis
tortion (as in rolled bars, plates, etc.) or by both or by machining (causing plastic
distortion in a surface layer under the pressure of the cutting tool).
CHAPTER 7

MECHANISM OF PLASTIC DEFORMATION


IN THE GRAIN STRUCTURE
The elastic deformation of crystalline materials under the action of
external forces is the result of distortion of the crystal lattice in which
the atoms, ions, or molecules are arranged. Likewise, the volume change
and elasticity, with respect to change in shape of vitreous solids and the
volume changes of liquids due to pressure, are the consequences of
the changes in the distances between the molecules. If we disregard the
elastic deformation, which at usual pressures up to several hundred or
thousand atmospheres is small in comparison with the possible per
manent (plastic) changes in shape, there remain for the explanation of
plastic flow in solid materials only the permanent displacements in the
relative positions of the elements or atoms in the crystal lattices.
The phenomena occurring during the plastic flow of the crystalline
solids are best studied by means of test pieces consisting of single crystals
of the size used as specimens for testing of materials.1
Although the mechanism of plastic deformation in the crystal lattice
of solid materials is not a simple one, some typical phenomena known to
occur in combination with plastic flow may be mentioned. The most
important of these elementary phenomena are the following:
1. Slip. Parallel displacements ("translations") of the elements of
the crystal lattice.
2. Formation of twins. Shifting as a whole, of a part of a crystal, to a
second position, symmetrical with respect to certain planes in the lattice
of the remaining part of the crystal.
3. Deviations from regular positions of atoms and their thermal agitation
in lattice cause motion of dislocations under stress.
4. Breakdown of structure, occurring often quite gradually under an
increasing load. Displacements of crystal grains accompanied by a par
tial destruction of the cohesion. This results in a gradual loosening or
tearing apart of the structure under increasing stress.
The phenomenon of slip or translation consists of a parallel sliding of

* Regarding the mechanical properties and the plastic


deformations of metal single
crystals which have been the subject of numerous valuable researches in recent years
see references on pp. 50-52.
49
50 THEORY OF FLOW AND FRACTURE OF SOLIDS

Fig. 7-1. Copper. Initially polished surface of specimen showing numerous slip bands
in the crystal grains, after plastic deformation. (Magnification about 100.)

parts of acrystal along planes, relative to one other, for distances which
are perhaps many thousand times the distance between the atoms in the
lattice. These displacements usually occur along one or more crystallo-
graphically definite planes in the crystal; moreover the directions of
sliding coincide with certain crystallographically definite straight lines in
the space lattice. Sliding in the grains often occurs in numerous parallel
planes, the number of planes increasing with the increasing stress. This
type of deformation often is evidenced by numerous parallel strips or
markings well observable under the microscope on the surface of a metal
crystal, or in the crystal grains of a deformed metal which has been
polished before it is subjected to a plastic distortion.
The laws of the plastic distortion in single metallic crystals for most
of the more important metallic elements and for several alloys have been
fully described by E. N. da C. Andrade,1 Constance F. Elam (Mrs. G. H.

Andrade, C, and P. J. Hctchinos, The Mechanical Behavior of Single Crystals


1

of Mercury, Proc. Roy. Soc. London, sor. A, vol. 148, pp. 120-146, 1935. Also Inter
national Conference on Physics, London, "The Solid State of Matter," Vol. II, p. 173,
Cambridge, University Press, London, 1934.
MECHANISM OF FLOW IX GRAIN STRUCTURE 51

Fig. 7-2. Copper. The deformed surface of a highly polished copper piece showing numer
ous slip bands in the crystal grains. The deformation was produced by the impression of a
punch. (Magnification about 100.)

Tipper),1 M. Polanyi, E. Schmid,2 Sir Geoffrey Taylor,3 and others." In


permanently extended single metallic crystals the slip planes are fre
quently crowded at regular intervals rather than distributed evenly over
1 "The Distortion of Metal Crystals"
(Oxford Engineering Science Series), Oxford
University Press, New York, 1935, 182 pp.
•Schmid, E., and W. Boas, "Kristallplastizitaet" (Struktur und Eigenschaften
der Materie, ser. 17), Verlag Julius Springer, Berlin, 1935, 373 pp. The books by
Elam and by Schmid and Boas contain the most complete description of the modes in
which the crystals with face- or space-centered and with hexagonal lattices deform
and their general laws of deformation and fracture by cleavage or shear.
3 Taylor, G. I., and C. F. Ela.vi,
Proc. Roy. Soc. London: The Distortion of an
Aluminum Crystal during a Tensile Test, Bakerian lecture, ser. A, vol. 102, p. 643,
1923; The Plastic Extension and Fracture of Aluminum Crystals, vol. 108, p. 28, 1925;
ser. A, vol. 112, p. 337, 1926 (distortion of iron crystals). Taylor, G. I., and W. S.
Farren, The Distortion of Crystals of Aluminum under Compression, Proc. Roy.
Soc. London, Part I, vol. Ill, p. 529, 1926; Part II, vol. 116, p. 16, 1927. Taylor,
G. I., Proc. Roy. Soc. London: The Mechanism of Plastic Deformation of Crystals, Part
I (Theoretical), Part II (Comparison with Observations), vol. 145, p. 362, 1934. See
also Taylor, G.
I.,

Proc. 2d Intern. Congr. Applied Mechanics, p. 46, Zurich, 1926.


Seitz, F., "Physics of Metals," already quoted and a paper by F. Seitz and T. A.
*

Read, "Theory of Plastic Properties of Solids," ./. Applied Phys., vol. 12, pp. 100-118,
52 THEORY OF FLOW AND FRACTURE OF SOLIDS

the length of specimens that have been pulled, and these regions are
separated by layers of apparently undistorted metal. A crystal which
originally was a circular cylinder
or a prism with a square cross
section deforms to a flat elliptic
cylinder or prism. The plane sec
tions which coincide with the
planes of slip remain congruent
ellipses or rhomboids and are tilted
toward the direction in which the
crystal is pulled. A deformed
metal crystal extended by C. F.
Elam is reproduced in Fig. 7-4.
^£31
Metals with the face-centered cubic
Flo. 7-3. Distorted .surface of an initially
highly polished copper specimen which has lattice (aluminum, copper, etc.),
been severely deformed. The slip bands for example, slip along the octahe
can hardly be detected. (Magnification
about 70.) dral planes in the direction of an
edge of the octahedron (the center
points of the six cubic planes define the corners of the octahedron).
Taylor and Elam found that an aluminum single crystal deforms accord
ing to the parabolic strain hardening function

r = c y/y
where r is the component of the shearing stress in the operating plane of
slip in the direction of the slip defining the unit shear 7. (The formula is

Fio. 7-4. Extended copper-aluminum single crystal. (Reproduced from


" Distortion
of
Metal Crystal*," by C. F. Elam, see p. 51n).

not valid for values approaching 7=0 and t = 0 because the crystal
starts to deform at a small but finite value of t. In crystals having a
face-centered cubic lattice "double slip" on two octahedral planes may
occur simultaneously also.)

170-186, 470-486, 538-554, 1941. Barrett, C. S., "Structure of Metals— Crystal-


lographic Methods, Principles, and Data," McGraw-Hill Book Company, Inc., New
York, 1943, 567 pp.
MECHANISM OF FLOW IN GRAIN STRUCTURE 53

By means of the electron microscope it has been observed1 on small single


crystals of a 99.99 per cent aluminum under increasing stress that one part of the
crystal slips relative to the other part in what appears to be a single plane by a distance
of about 2,000 angstrom units (1 angstrom unit A = 10~" cm). After the move
ment in this plane ceases a slip begins anew on a parallel plane 200 A apart and further
slips follow in succession at similar distances until a package of planes 200 A apart has
formed. In their experiments with single crystals of mercury Andrade and Hutchings8
found that the glide planes were 0.005 cm apart and that there were approximately
200,000 atomic planes between them. Mercury (which crystallizes in the rhombo-
hedral system) glides on rhombohedral planes in the direction of the short diagonal,
but it distorts frequently also by "twinning" (see page 61). At — 180°C the plastic
distortion proceeds more readily by twinning than at — 50°C. The direction of slip
in a iron (having the body-centered cubic lattice) is along a space diagonal in the cube,
but the slip planes can take several orientations having simple rational ratios between
the intercepts in which these planes intersect the three crystal axes (e.g., planes con
taining two opposite edges of the cube and two face diagonals may develop into glide
planes). At — 185°C a iron according to Schmid and Fahrenhorst3 deforms very
little by slip, but clearly by twinning. Fracture in an a iron crystal at this low tem
perature either occurs by cleavage along a cubic plane or the crystal may be pulled
out to a thin wedge along which it breaks.'

Herbert J. Gough' and Herbert F. Moore" in their fundamental studies


of the fatigue failure in metals under oscillating stresses found the sur
faces of the crystal grains in the neighborhood of fatigue cracks in broken
specimens frequently densely covered by parallel slip lines; and they
concluded from these observations that, since the orientation of fatigue
cracks favored the direction parallel to the slip bands, the more important
types of failures by fatigue in metals are related to or are a consequence
of slip within the crystal grains. (Two examples of the development of
1 Heidenreich, R. D., and W. Shockley, a note in J. Applied Phys., vol. 18,
p. 1029. Also Heidenreich and V. G. Peck, Fine Structure of Metallic Surfaces
with the Electron Microscope, J. Applied Phys., vol. 14, pp. 23-29, 312-320, 1943;
Phys. Rev., vol. 62, p. 292, 1942.
* hoc. cit.
» Z. Physik, vol. 78, p. 383, 1932.
* The mechanism of slip in pure iron and in silicon iron has been studied by many

investigators. Taylor and Elam found frequently "wavy" slip lines, explaining
them as the traces of movements by slip along the space diagonals of the body-cen
tered lattice of a iron of bundles of pencil-like small bodies. 0. S. Barrett, G.
Ansel, and R. F. Mehl (Trans. ASM, vol. 25, pp. 702-733, 1937) observed straight
slip lines in pure iron crystals and that slip can occur in iron in three preferred planes
of the crystal.
s "Crystalline Structure in Relation to Failure of Metals Especially by Fatigue,"

Edgar Marburg lecture, 1933 (ASTM meeting, Chicago).


8
Moore, H. F., and T. Ver, "A Study of Slip Lines, Strain Lines and Cracks in
Metals under Repeated Stress." Moore, H. F., B. B. Betty, and C. W. Dom.ins,
"The Creep and Fracture of Lead and Lead Alloys," Univ. Illinois Eng. Expt. Sta.
Bull. 208, 1930; Bull. 272, 1935.
54 THEORY OF FLOW AND FRACTURE OF SOLIDS

fatigue cracks taken from two papers by H. F. Moore and his collabo
rators are reproduced in the micrographs Figs. 7-5 and 7-6. In the first
the crack followed the slip lines, in the second the grain boundaries; the
first refers to an example of a strong, the second to a very weak material.)
Although the phenomenon of discontinuous slip along crystallographi-
cally determined planes must be a part of the mechanism of the strain
hardening of those metals that deform by slip, it must be remarked that
the process of simple translations by multiples of the lattice distances

Flu. 7-5. Fatigue crack following slip linos Fig. 7-6. Fatigue crack along grain bound
in monel metal. (Magnification 370.) (//. aries in antimony-lead alloy. (Magnifica
F. Moore and T. Ver.) tion 37.) (H. F." Moore, B. B. Betty, C. W.
Dollins.)

which would leave the ideal crystal in an elastically unstrained condition


after a specimen was deformed and unloaded still cannot explain strain
hardening in metals for several reasons. If slip occurred at once by
multiples of the atomic distances in all points of a slip plane, the interior
of a crystal would remain unchanged. It would be difficult to see why
the shearing stress r should increase with the unit shears y according to
the observed strain hardening function t = f(y). It is well known, on
the contrary, that specimens of metals after a nonuniform plastic dis
tortion in which they have not been strained in all parts of the body by
the same amount (as by plastic bending or torsion in bars) upon removal
of the external load contain internal stresses. A system of residual or
inherent stresses acts, and elastic energy is stored in them. The same
must be true in the polycrystalline structure of a metal on a microscopic
scale in any region which has uniformly been permanently deformed (as
by simple tension), because some grains have been strained more, others
less, with the result that, after removal of the external load, systems of
residual stresses remain in the grains and elastic energy is stored in them.
MECHANISM OF FLOW IN GRAIN STRUCTURE 55

Near the grain boundaries (in deformed single crystals near their unde-
formed heads) the packages of the slip bands are clearly seen bent to a
considerable degree.
Carpenter and Elam1 in 1921 found that the deformation at the bound
ary of the crystallites in coarse-grained polycrystalline aluminum was
much less than in the center of the crystals. W. Boas and M. E. Har-
greaves2 made careful observations on small flat tensile specimens of
a 99.15 per cent purity aluminum electrolytically polished consisting
of a few large grains of 1 to 3 cm mean diameter and found that both the
unit elongation and the Vickers hardness within a few large crystals
measured along the axis of the specimen systematically deviated
from their mean values after the aluminum bars had been stretched
permanently in tension by 5 and subsequently by 17 per cent more.
The unit elongation in one single large grain of aluminum varied between
the values of 3 and 6 per cent and between 10 and 18 per cent local strain,
respectively. They concluded that, during plastic deformation of an
aggregate of crystals, there are wide differences in the amount of deforma
tion and work hardening from grain to grain, furthermore that the defor
mation is inhomogeneous within the grains because of their interaction.
This then clearly shows that the mechanism of simple or double slip
cannot account for the observed local variations in the strains from
grain to grain and within the crystals themselves and that a more complex
process contributes to the elementary mechanism of plastic distortion
in a crystal which is responsible for the mean permanent strain in a work
hardened aggregate of grains.
H. Hort,3 G. I. Taylor, W. S. Farren, and H. Quinney4 have shown by
calorimetric measurements of the heat evolved during the perma
nent stretching of metallic bars that the mechanical work done is not
entirely converted in heat. A measurable portion (about 10 per cent or
less) of the work becomes latent in the form of elastic stored energy.
When a metal strain hardens, thus small regions in it must be strained
elastically in some manner. It is remarkable in this respect that recrys-
tallization starts easily from the grain boundaries, namely, from those
regions which must have been strained elastically the most when the
metal was cold worked, on account of the disturbed conditions in

1 Proc. Roy. Soc. London, ser. A, vol. 100, pp. 329-353, 1921.
* On the Inhomogeneity of Plastic Deformation in the Crystals of an Aggregate,
Proc. Roy. Soc. London, ser. A, vol. 193, pp. 89-97, 1948.
* Waermevorgaenge beim Laengen von Metallen, Forschungsarb. Ver. dent. Ing.,
No. 41, 1907, 53 pp.
* Proc. Roy. Soc. London, ser. A, vol. 107, p. 422, 1925; vol. 143, p. 307, 1934; vol. 163,

p. 157, 1937.
56 THEORY OF FLOW AND FRACTURE OF SOLIDS

the lattice near the boundaries. The microscopic systems of residual


stresses located near the grain boundaries facilitate the new growth of
grains originating from nuclei in the latter regions when a metal is
1
reheated.
The question whether the lattice in the crystals of a fine-grained metal undergoes
some changes in those polycrystalline metals which are known to have a pronounced
yield point or remains unchanged after plastic deformation has been investigated by
S. L. Smith and W. A. Wood* at the National Physical Laboratory, Teddington,
England. They tested small flat specimens of pure iron (99.95 % Fe) and of normal
ized mild steel (0.1 % C) in tension and simultaneously observed the lattice distances
in the grains of these metals when the specimens started to yield and during their
subsequent plastic distortion until the maximum load was reached by means of the
X-ray reflection method. By observing the small changes of the lattice distances due
to the load and the plastic distortion in a fraction of the crystal grains (favorably
oriented with respect to the incident X-ray beam) in which certain atomic planes in
the body-centered cubic lattice of the a iron deviated by a certain small angle from
the plane perpendicular to the direction of pulling, they were able to state that the
iron lattice first deformed perfectly in an elastic and reversible manner corresponding
to the elastic range of the strains below the yield point. However, after the yield
point was reached, they observed in the direction perpendicular to the pull a certain
very small expansion of the lattice distances which remained stationary after the load
dropped from the upper to the lower yield point which is a characteristic property of
the behavior of steel. This expansion was a permanent one and remained after the
specimens were unloaded but increased slightly further under increasing flow stresses.
The permanent expansion in the lattice distances was estimated to be 0.03 per cent
at the yield point in the directions perpendicular and parallel to the tension stress, a
figure which appears to be perhaps too high, since it would imply a decrease of the
density of the metal by nearly a thousandth.'

The discrepancy between the comparatively very small values of the


shearing stresses required to initiate the plastic distortion in single
metallic crystals which were experimentally determined and the
1 Evidence that metal crystals may be deformed clastically in very thin lamellae or
break up in small blocks of lamellae which are slightly tilted with respect to the slip
planes is mentioned on p. 200. The presence of locally distributed elastic strains in a
metal (after it has been cold worked) is disclosed by the "asterism" (diffusely
elongated spots) in the X-ray diffraction patterns of metals. Seitz, F., loc. cit.,
p. 80; Barrett, C. S., "Structure of Metals," p. 349 (Metallurgical Engineering
Series), McGraw-Hill Book Company, Inc., New York, 1943, 567 pp.
* A Stress-Strain
Curve for the Atomic Lattice of Iron, Proc. Roy. Soc. London, ser.
A, vol. 178, pp. 93-106, 1941; A Stress-Strain Curve for the Atomic Lattice of Mild
Steel, and the Physical Significance of the Yield Point of a Metal, Proc. Roy. Soc.
London, ser. A, vol. 178, pp. 45-460, 1941; Proc. Roy. Soc. London, vol. 182, p. 404,
1944. Also Wood, W. A., The Behavior of the Lattice of Polycrystalline Iron in
Tension, Proc. Roy. Soc. London, vol. 192, pp. 218-231, 1948.
3 A decrease in density of severely cold
worked metals is known to occur, but it has
heen attributed to the formation of voids between the lamellae of elastically strained
packages slipping relative to each other in the deformed crystallites.
MECHANISM OF FLOW IN GRAIN STRUCTURE 57

high values of the shearing stresses causing a translation in a crystal


which are computed from the atomic theory has inspired investigators to
discuss realistic models of the arrangement of the atoms in solids as
distinguished from the "ideal" crystal lattices. Two of these
mechanical models of solids will be briefly described here; a few additional
modifications of such models have been discussed by physicists.

Consider according to L. Prandtl a row of material points mi, mi, . . . (represent


ing a string of atoms) having one degree of freedom for the sake of simplicity aligned
on a frictionless rod (Fig. 7-7; the rod is not shown) and bound to solid posts Si, Si,
... by means of equal elastic springs. This row of masses is supposed to represent,

Fig. 7-7. Prandtl 's model.

for example, the atoms in the surface of one crystallite nearest to a second grain. If
no action were exerted from the neighboring crystallite, the masses mi, mt, . . .
would take their rest positions midway between the supports Si, St, . . . We must
assume, however, that the atoms in the surface of the second crystallite exert certain
forces on the masses mi, mj, . . . and again, for simplicity, we shall suppose that they
are of an elastic nature similar to those of the springs. We may visualize the field of
these forces by a sinusoidal wave as drawn above line AG parallel to the line support
ing the masses mi, ms, ... in Fig. 7-7. It is understood, since only one degree of
freedom of motion is to be considered, that the ordinates of the sine curve ABC . . .
represent components of force which act in the direction of the row of masses mi, m2,
... or the horizontal components of the force field (positive ordinates plotted above
line AG represent force components driving the masses in Fig. 7-7 from right to left) ;
also that the wave length of the cosine curve must not necessarily coincide with the
spacing between the rest positions of mi and ms, ...; it may not be much different,
however, since curve ABC . . . represents a force field emanating from a regular
lattice of the same kind of atoms from the neighbor crystallite. When a state of
stress acts in the body (to which both crystallites belong) this will have the effect that
under an increase of stress the row of atoms mi, ms, ... is slightly displaced past the
sinusoidal force field and in its direction. The field will move the masses mi, m;, . . .
from their former rest positions Ri, Rt, . . . to new equilibrium positions, such as
those shown in the lower part of Fig. 7-7 by the points E\, 2?s, . . . Obviously the
58 THEORY OF FLOW AND FRACTURE OF SOLIDS

positions of the points E\, Ei, . . . are determined by the condition that in these
points the elastic forces of the springs must become equal and of opposite sign to the
forces exerted by the neighbor crystallite which were represented by the ordinates of
curve ABC .... These positions are easily found by drawing the straight lines
RiB, R2F, . . . , etc., from the rest positions of the masses toi, mi, . . . under equal
slopes, assuming that their ordinates represent the elastic forces with their signs
reversed. Their points of intersection B, F, . . . with the cosine curve determine
those new equilibrium positions A'i, Ei, ... If the elastic springs are very stiff, the
straight lines R\B, . . . must be steeply inclined with respect to AG, and if the atoms
are more loosely bound to their neighbors, they must have smaller slopes.
A little thought may convince us that there will always be stable positions of equi
librium, if the straight lines are steeper than the maximum slope of curve ABCDF
... is. Figure 7-7 illustrates just the other case, in which a weaker spring constant
was assumed and a slope chosen, under which a tangent line may be drawn to the sine
curve, such as the one passing through point C and point D. In the latter case (when
the elastic bonds connecting the atoms mi, m2, . . . with the columns Si, Si, . . .
are softened) one sees easily that now unstable equilibrium positions for the mi, ms,
. . . also exist. (An equilibrium will become unstable if the difference between the
spring and field force ceases to be a "restoring" force for a small deviation of the mass
from its new equilibrium position, i.e., if this force points away from the latter.) In
the relative position of frame St, Si, 1S3, . . . and field ABC . . . shown in Fig. 7-7
both equilibrium positions of mi and mi at the points E\ and Ei are stable. Suppose,
however, that the force field is slowly displaced to the left relative to the posts Si,
Si, St, . . . Point B would move upon the sine curve toward C. At the instant when
B coincides with C the equilibrium of mass mi ceases to be stable (for a small deviation
of mi to the right) and mass vi\ will be driven in the new position marked by C\ around
which it will oscillate violently until its kinetic energy is dissipated as heat by damping.
The same happens to mass mi if the field is displaced to the right. When point F
coincides with point D, mi will jump to point D\. If the picture of Figure 7-7 is
extended on both sides so that it includes many masses m, those whose positions are
determined by points situated within the arcs CD of the sine waves will all suddenly
be caught in unstable positions in which they cannot remain. This model represent
ing greatly simplified conditions near a grain boundary or in the neighborhood of an
internal imperfection in a solid on the atomic scale was proposed by L. Prandtl in
1913 and served him1 to compute the relation connecting stress and strain during
repeated loading and unloading cycles (the law for elastic hysteresis, see page 27)
and the logarithmic dependence of the yield stresses on the plastic rates of strain
(see page 21).*
Instead of displacing the force-field curve relative to the frame carrying the masses
m we may move the latter with respect to the former and may plot the force F estab

lishing equilibrium for the mass m, (the ordinate E\B) in function of the relative
displacement x. A curve will be obtained F = F(x) as shown on the right of Fig. 7-7
by curve ACCiG and if the direction of relative motion is reversed by curve GDDiA.

1 Described in Ein Gedankenmodell zur kinetischen Theorie der festen Koerper,


Z. angew. Math. Mechanik, vol. 8, pp. 85-106, 1928.
8 He also included formulas expressing the decrease of the viscosity of fluids
with
the temperature well known from tests on mineral oils and the rapid increase of vis
cosity with increase of the hydrostatic pressure which was observed by P. W. Bridgman
(1926).
MECHANISM OF FLOW IN GRAIN STRUCTURE 59

For an oscillation between the values xd S x g xc a closed curve for F(x) will be
obtained representing an "elementary" hysteresis loop for one single atom. Prandtl
considers the effect of the absolute temperature T on these and related phenomena.
The probability that a jump of an atom such as CC, or DD\ in Fig. 7-7 would occur
because of the thermal agitation according to kinetic theory is expressed by e~T^T
(e base of natural logarithm, To a constant). An
atom can leave a stable position prematurely in
which it was near the crest of the sinusoidal force
curve or can climb up along it (from a point near the
crest of the wave to the left of point C in Fig. 7-7) by
doing the work required against the difference of the
elastic and field forces (the shaded area to the left of
point 0) or against this barrier of potential energy u
if its kinetic energy of thermal agitation just exceeds
this amount. The probability that this excess of
energy u would become available was assumed equal
to e~n/ut, where the quantity u0 is proportional to the
absolute temperature T. The elementary process of
changes in position of atoms in the crystal lattice
from their regular positions through unstable to new
stable ones provided the means for deriving the func
tional properties of cyclic stress-strain curves (elastic
hysteresis), the expressions for the elastic after (time) Flo. 7-8. Sir Geoffrey Tay
lor, F.R.8., Cambridge, Eng
effect and the logarithmic speed law <r <r\ + <r, land.
In (t>/tio) connecting the plastic rates of strain v with the
yield stress <r(au<ro,t'o const) in metals, including their dependence on the temperature.
Although strain hardening was not considered by Prandtl, a mechanical model for
it was proposed by Sir Geoffrey Taylor1 in 1934 in his two-dimensional mechanism of
the slip in crystals. In this model certain deviations are considered frozn the regular
positions of the atoms, and these disturbances are shown to move under a shearing
stress t through the undisturbed parts of the crystal, causing one portion in the crystal
to be displaced by one atomic spacing (a slip by one atom spacing) with respect to
the other portion. Taylor assumes that misfits exist in parallel planes of atoms at
periodic intervals and supposes that the rows of atoms on one side of these planes are
slightly contracted while the rows on the other side are slightly extended. This
amounts to assuming in the lattice one more atom above than below the plane A A
over a certain length (Fig. 7-9). The forces exerted by the atoms of these two rows
will cause a certain distribution of the potential energy in the plane A A. In a normal
lattice (on both sides of the plane) this curve is a cosine function as in Prandtl's model,
and if atoms were placed in the troughs of the cosine wave, they would stay there in
stable equilibrium. In the case, however, of one row having one atom less than the
other row the energy curve is symmetrically distorted with respect to the center of
the length and the amplitudes of the energy potential waves are seen to decrease
toward the center of symmetry (Fig. 7-9). If the crystal is strained elastically by a
shearing stress t the two atom rows are slightly displaced in the tangential direction
relative to each other, and this causes the curve of the potential energy in plane A A
to become unsymmetrical with its shallowest trough in the center of the disturbed
1 The Mechanism of Plastic Deformation of Crystals, Tart I, theoretical, and
Part II, comparison with observation; Proc. Roy. Soc. London, vol. 145, pp. 362, 388-
404. 1934.
60 THEORY OF FLOW AND FRACTURE OF SOLIDS

length and one next deeper trough on one side of the former (Fig. 7-9). If atoms were
placed in the troughs with exception of the shallowest one and the thermal agitation
(vibration) of the atoms were permitted to act at a moderately low temperature at
which the mean kinetic energy corresponding to these vibrations just equal to the

is
height of the hill to the left of the atom nearest to the vacant place, this atom would
be thrown out first from its least stable
position of equilibrium into the vacant
hollow to its left because was separated

it
by the lowest potential barrier (crest of
wave) from the next stable position.
The hollow defining its former position
would thus become vacant. The action
of the forces emanating from the middle
row in plane AA on the two neighbor
rows has not been considered which must
cause rearrangement of the positions

a
of the atoms in these two rows. Taylor
assumes that after this rearrangement
the center (the new vacant hollow) has
T been displaced to the right, including the
VACANT I
condensed and rarefied regions in the
ioj—
-cj--

--<{
<^ lattice. The final effect of the jump of
one atom in the plane to the left

A
A
will be that the center of the entire dis
turbance in the lattice has been moved
by one atom spacing to the right in the
arrangement of Fig. 7-9. If the com
pressed and extended rows were assumed
below and above plane A, respectively,
A
and the direction of the shear stress t
if

Fio. 7-9. Potential energy in plane A. remained the same as in Fig. 7-9, one sees
A

(a) When no external stress is applied. that the center of this second disturbance
(b) When lattice is subjected to a shear which the mirror image of the first one
is

stress. Theory of DUlocationa," by G.


I.
("

with respect to plane was moved by


A
A

Taylor.)
one atom spacing to the left. It will be
noted that the passage of both types of "dislocation" (Taylor calls the first a
positive, the second a negative dislocation) across crystal produces the same effect,
a

namely, a shift of the block above plane AA by one atom spacing to the right relative
to the block below A. It also of interest to note that the state of stress around the
A

is

center of a dislocation (with the exception of its immediate neighborhood) may be


computed by making use of a known solution of the theory of elasticity and that, in
the wake of such disturbance after a slip by one atom spacing has occurred, new
a
X
a

shear stress induced around its center x = = 0:


r'
is

_ G\ x
,
7

« i2+J2
the modulus of rigidity) which decreases inversely as the distance
G
is

(where

= \/x2 +
I/'
T

measured from the center of the disturbance. The direction of created by a positive
t'
MECHANISM OF FLOW IN GRAIN STRUCTURE 61

dislocation is such that r' will increase the shear stress t to the right of the center, i.e.,
when x is positive (in Fig. 7-9). Therefore, a second positive dislocation situated in
a plane parallel to plane A A to the right of the first one (i > 0) will be repelled by
the first one, and vice versa when t = 0. Two positive and similarly two negative
centers will repel each other, but one positive and one negative dislocation will be
attracted to each other until they take a certain relative position dependent on the
value of r.
Various multiple systems or regular configurations of positive and negative disloca
tions might be considered on parallel planes at regular intervals. Taylor visualizes
the process of plastic deformation by assuming that slipping in the crystal begins at
centers distributed at random because of thermal agitation whereby positive and
negative dislocations are separated. These move across the lattice planes a certain
average distance under the shearing stress t until they are stopped at surfaces of
misfit. The mean distance which dislocations can travel represents an essential
parameter in this theory. A number of observations appear to support the existence
of such surfaces of misfit at periodic intervals in crystals. It is thus supposed that
sliding occurs over limited lengths. This theory leads to a parabolic relationship
between the shear stress t and the plastic unit shear y [see page 52]. It also explains
a spatial fluctuation in the values of the shear stress t + t' caused by the high values
of the shear stresses t' originating from the centers of the dislocations which were
stopped at the internal surfaces of misfit and demonstrates that certain amounts of
elastic energy are stored in these regions.1

The second kind of phenomenon occurring during plastic deformation


of a crystal, known as crystal twinning, is a shifting
of the position of the lattice in a part of a crystal into
a second position, so that the second part is in a sym
metrical position relative to the first part and some
plane of symmetry. An example is exhibited by the
deformation of crystals of calcite. These crystals tion of twins in "a cat-
split if loaded in a suitable way,2 in the manner cite crystal. (Ac-
COT tn° vo
shown in Fig. 7-10. This kind of deformation, also
designated by Mtigge3 as "shear," has been observed in the crystal grains
of naturally or artificially deformed marble4 and in certain metals.
1 Sir Lawrence Bragg recently proposed a model with which the movements in a
regular arrangement of small bubbles of air of perfectly equal size floating on a soap
solution may be seen which are caused by a local misfit of the bubbles in one row (an
accidentally larger bubble embedded between two neighboring rows) when the layers
of bubbles are strained by shearing forces. This model served him for illustrating
the movements of dislocations in crystal lattices according to Taylor and other
phenomena. (Braoo, L., and J. F. Nye, A Dynamical Model of a Crystal Structure,
Proc. Roy. Soc. London, ser. A, vol. 190, pp. 474-481, September, 1947.)
* Niooli, "Lehrbuch der Mineralogie,"
p. 176, Verlagsbuchhandlung Gebrilder
Borntraeger, Berlin, 1920.
* " Kristallphysik, Handbuch der Naturwissenschaften," Vol. 5, p. 1135, 1914.
4 See the tests under combined axial and lateral pressure, made with marble
test
pieces of cylindrical shape by T. v. Karman, Mitt. Forschungsarb. Ver. dent. Ing.,
No. 118, Berlin, 1913.
62 THEORY OF FLOW A<\D FRACTURE OF SOLIDS

As has already been mentioned, the shear stress, necessary to produce


the plastic state, is a definite function of the temperature. Hence in the
mechanism of plastic deformation of solid materials under high tempera
ture, the thermal vibrations of the atoms must play an important part.
The capacity for forming slip planes in the crystals is considerably
increased by heating. Thus near the melting temperature the formation
of slip planes may be brought about by very small shear stresses.
Among noteworthy contributions to the theory of slip in crystal lattices
which preceded the work by Prandtl and Taylor, reference is made to
papers by R. Becker, A Smekal, and E. Orowan.1 We may accord
ing to R. Becker conceive of certain layers in the crystal whose boundary
planes coincide with the slip planes of the crystal lattice. These layers
are bounded by rough surfaces. The roughnesses stand out like teeth
or bumps and fit in corresponding depressions in the bounding layers.
These correspond to the force fields of the atoms. For example, in
single crystals of copper, whose atoms form a face-centered cubic lat
tice, in which the octahedron planes act as slip planes, and the edges
of the octahedron represent the direction of slip, the middle points of these
bumps must lie upon a net composed of a mesh of equilateral triangles.
The forces emanating from the atoms hold the layers together. As soon
as the laws obeyed by the interatomic forces are known, it is possible to
calculate the shearing force S necessary to cause a displacement of these
layers sufficient to bring these bumps opposite each other. This force <S
would then produce failure by slipping. This theory of action within
crystals was extended by R. Becker to include the effects of the thermal
agitation of the atoms. In general terms this may be explained as
follows: If we assume that the shearing force in the crystal has a value
but little smaller than the force S necessary to produce permanent slip,
it follows that there will be elastic deformations only in a lattice with the
atoms at rest. However, if we consider the thermal vibrations of the
atoms due to heat, the picture changes. Since a variable shearing force s
is exerted by the vibrating atoms it follows under these conditions that it
varies periodically about a mean value S„. It is thus possible that in
certain planes the force s may for an instant become greater than S.
When this occurs, as mentioned above, slip in the layers of the crystal
follows, where s > S. The plasticity of the crystal, according to this
theory, is affected by the temperature; the higher the temperature the
1 Becker, R., tlber Plastizitat, Vcrfostigung und Rekristallisation, Z. tech. Physik.
No. 7, p. 547, 1926; Smekal, A., Zur Molekulartheorie der Festigkeit und der Ver-
fcstigung, ibid., No. 11, 1926; Am. Akad. Wins. Wien, Dee. 2, 1926, Jan. 27, Mar. 17,
1927; Orowan, E., Zur Kristallplastizitat, three papers in Z. Physik, vol. 89, pp. 605,
'
614, 634, 1934. .
MECHANISM OF FLOW IN GRAIN STRUCTURE 03

more frequently will the slip layers occur. The velocity of plastic defor
mation would depend on the number of planes in which sliding occurs.
Although the theory of Becker was an important step in the direction
of considering the effect of temperature on the phenomenon of slip in
crystals, a formula which he derived for the dependence of the shear
stress on the plastic rate of shear is not veri
fied by experiments and will not be quoted
here. E. Orowan1 suggested interesting
ideas for improvements of this theory by
assuming a misfit in a slip plane whose region
spreads over a circular area around the center
of the disturbance in the slip plane2 and which
was similar to the dislocations described by
Taylor, but he did not elaborate on a detailed
analysis of these imperfections.
On the basis of the conception of small
flaws of Griffith,3 A. Smekal assumed that
crystals (even the most perfect specimens),
as well as amorphous solids, are weakened
by countless small flaws. By means of in
genious experiments on small bars of rock
salt, plastically bent (by coloring them by
means of exposure to radium radiation),
he was able to show that in the plastically
J^JJi. RiES
deformed part of rock-salt crystals the loosening up in the lattice was
greater than in the unstrained portion. Such changes of the microstruc-
ture of ductile metals by means of severe cold working are well known to
engineers. The metal becomes "more brittle" after severe cold working.
On the other hand Polanyi and E. Schmid have shown,4 that in metal
single crystals plasticity due to ordinary slip can be observed at the lowest
possible temperatures (liquid helium). Hence one main part of plasticity
due to ordinary slip must be attributed to a mechanism which is not
affected by the thermal agitation of the atoms.
F. Zwicky* has raised doubts against the assumption that the presence

1 Loc. cit.
' hoc. cit., p. 639.
• See the theory of fracture by A. Griffith, Chap. 15, p. 196.
4 Natunixissenschaflen, vol. 17, p. 301, 1929; rf. also several papers by E. Schmid in
Z. Physik, 1929 and 1930.
* The Imperfections of Crystals, Proc. Nat.
Acad. Sci. U.S., vol. 15, p. 253, 1929.
On Mosaic Crystals, ibid., vol. 15, p. 816, 1929; also ibid., vol. 16, p. 211, 1930. See
also Helv. Phys. Acta, vol. 3, p. 269, Zurich, 1930.
64 THEORY OF FLOW AND FRACTURE OF SOLIDS

of imperfections (minute cracks or flaws) in crystals should exert the large


effect on certain mechanical properties of the crystals (such as strength,
the plastic limit) as had been supposed.
In the solid state of matter, according to Smekal and Zwicky, two
kinds of properties have to be distinguished. Certain physical properties
of the crystals are known to be structure insensitive, while others are
structure sensitive. To the first group of physical properties of the crystal
lattices belong density, specific heat, elasticity (compressibility), thermal
expansion, and others; to the second belong mechanical strength, the
limit of plasticity, dielectric strength, certain optical properties, and
. others. The first kind of properties seem to have about the same values
for single crystals as for the polycrystalline material of the same chemical
constitution. The latter group seems to be affected more than the former
by impurities, by a previous deformation, and by temperature (anneal
ing, tempering).1
According to Zwicky, there are two reasons against the validity of the
assumption that minute cracks and flaws are the main cause of the lower
ing of the mechanical "molecular" strength as deduced from atomic
theory to its actually observed low "technical" values.2
On the basis of these views, one would be led to expect that the behavior
of the real crystals would become more similar to that of the ideal crystals,
depending on the degree to which accidental disturbances are avoided
during the growing of the crystals. Observations, however, show that
the contrary is true. The second reason is that, if the lowering of the
strength is caused by a random distribution of minute cracks, the values
of the properties observed in actual tests would have to be distributed
according to rules of probability, while as a matter of fact they can be
reproduced within comparatively narrow limits.
To avoid these and other difficulties not mentioned here (which arise
from the above assumptions, according to Zwicky) small but finite
periodic changes consisting of slight variations in the distances within
1 Zwicky himself states that a sharp limit does not exist between these two groups of
properties. The elasticity constants or the thermal expansion coefficients of metals,
for example, are known to change with increasing temperature to a considerable
amount. The variation in these properties is perhaps of the order of 25 to 50 per cent
if the metal is heated from room to melting temperature. Possibly some of the
structure-insensitive properties change with rising temperature and become at ele
vated temperatures structure sensitive.
2
According to Polanyi and Zwicky and to the atomic theory, the molecular tensile
strength of, for example, a rock-salt crystal should be about a tenth of the mean value
of the elasticity moduli of this substance a = Z?/10 or about 2 to 4.104 kg/cm*, while
actually rock salt becomes plastic even at a stress of 20 kg/cm* or so and breaks at a
stress of perhaps 50 kg /cm*.
MECHANISM OF FLOW IN GRAIN STRUCTURE 65

the lattice must be assumed in the crystals. Upon the primary lattice
of the ideal crystal, as determined by X-ray analysis, secondary disturb
ances are superimposed, these consisting of small periodic variations of
the density or of the distances between the elements of the lattice. The
structure-insensitive properties are caused by the primary lattice, the
structure-sensitive properties by the secondary disturbances.
These investigations show that the ideal lattice is thermodynamically
less stable than the mosaic crystal, i.e., a crystal having slight periodic
variations (contractions) in the lattice. This secondary structure within
the primary lattice would cause effects such as lowering of strength,
plastic deformation, and others. To give
an idea of the order of magnitude of the
blocks caused by the secondary structure,
it may be said that the distance between
two contracted parts within the lattice of
a rock-salt crystal are thought to be per
haps of the order of twenty times the dis
tances between the elements in the undis
turbed lattice,' so that the cubic block of ,,_... .
riG. 7-12. Longitudinal section of
the mosaic crystal of rock Salt would con- a paraffin test piece after compres-

tain about 8.000 to 10,000 atoms. This showin8 destruction of


' slon 4test
structure.
checks with Smekal's previous estimates
of the size of the seconary blocks limited by the imperfections.
The phenomenon of a gradual breakdown of the structure with increas
ing stress consisting in countless relative movements of parts of crystal
grains to each other may also be observed on a larger scale. In materials
with a relatively loose microstructure, a kind of plasticity seems to appear
under increasing stress. Permanent deformations result, since the
cohesion under the increasing load is gradually destroyed. An extreme
case of this kind is illustrated by a conical-shaped body made of paraffin
and loaded under compression (Fig. 7-12). A longitudinal section of this
body shows that the structure is broken down or severely deformed, much
worse at certain points than at others. The parts having the destroyed
structure are recognized by the discoloration or bright color. Another
sample is shown in Fig. 7-13 exhibiting a remarkably regular system of
shear lines. Many brittle materials may probably behave in a similar
manner when stressed by axial compression or tension. The small
crystal grains loosen or break prematurely before any motion of transla
tion may occur. For example, in a compression test with a cast-zinc
test bar we may observe, simultaneously with the occurrence of slip
surfaces in the crystal grains, the appearance of countless fine hair cracks
which prematurely destroy the cohesion of the material.
66 THEORY OF FLOW AND FRACTURE OF SOLIDS

When discussing the chances of failing of a single grain in an aggregate


of crystallites of a certain class oriented with their axes at random rela
tive to the axes of principal stress, R. Boker1 considers that a grain in
general can fail in four different manners: (1) by simple slip (including
twinning) through a translation along parallel planes within its crystal
lattice; (2) if the limiting value of
the shearing stress causing plastic
slip by translation is comparatively
large, the grain may not be able to
deform in the preceding manner
unless some corners in the adjoining
grains are sheared off (are split by
shear). If tensile normal stresses
are present, a grain may have two
more chances for failing: (3) by
simple cleavage along a crystallo-
graphically favorably inclined plane
(intracrystalline cleavage) or (4) by
separation along the grain bound
ary. If the tensile strength of the
grain boundary is smaller at some
point in which the tensile stress is
larger than the stress under which
favorably inclined
one of the most

Fio. 7-13. Regular markings on face of Planes in the crystallite Can break
compressed prism of paraffin. by simple cleavage, the boundary
will split open at this point.
The various phenomena, above considered, characteristic of plastic
deformation, often occur in combination with each other. Hence it
may not be possible to separate them from one another. This explains
the great variations in the mechanical behavior of solids encountered
when they are brought into the plastic state.
We have mentioned that crystalline solids and especially the metals as
well as the amorphous solids under a sustained load at sufficiently high
temperatures deform continuously and creep. Although the phenomena
of slow creep of the technical metals cannot be treated in this volume, it
may be in order at least to mention here various changes which have been
observed in the grain structure of alloys and pure metals after they have
been deformed in long-time tensile tests at high temperatures. One of
the best known changes is a breakdown of the metal along the grain
boundaries, characteristic of creep, particularly when it reaches its final
1 Aachen, 1914, p. 31.
Dissertation,
MECHANISM OF FLOW IN GRAIN STRUCTURE 07

stage and fracture is imminent. An example is given in the views of two


tensile creep test bars reproduced in Figs. 15-15 and 15-16. In a micro
graph taken from an investigation on the creep of lead and lead alloys
by H. F. Moore1 which is reproduced in Fig. 7-14, one will notice that
fine microtome scratches, which were present on the surface of the speci
men before the creep tests were started, became broken lines because of
the distortion and rotation of the crystallites and furthermore that the
grain boundaries became distinctly
visible. These lines appeared
darker and thicker after creep had
occurred than they were in the
original condition of the metal. It
is well known that the fracture of
alloys in long-time tensile creep
tests is intercrystalline and follows
the boundaries of the grains in con
trast to the transcrystalline frac
tures of metals observed at a mod
erately elevated or at normal
temperature. The breakdown of
the grain boundaries seems already
at work while the metal creeps
slowly. It is claimed that crystals Fio. 7-14. Two per cent tin-lead alloy
of some strong and brittle materials after 1,100 hr at 1,000 lb/in.2 stress.
(Magnification ISO.) (//. F. Moore, B. B.
(quartz, precious minerals) behave Betty, C. W. Dollins.)
almost perfectly in an elastic man
ner at normal and probably also at not too high temperatures.
It has been definitely shown, on the other hand, that metallic single
crystals at comparatively low temperatures show the phenomenon of
creep. According to tests by B. Chalmers,2 single crystals of pure tin
under constant small tensile stresses start to elongate with an initial very
small rate of strain v which is proportional to the stress a = cv. These
rates of creep decrease gradually with time under a constant stress. If
the stress is raised above a certain value the tin crystals creep continu
ously under a constant stress; in this range the initial rates of creep v
and also the rates with which the crystal stretches after a longer time
has elapsed increase much faster than the stress a. If a is plotted
as a function of ti a curve is obtained which rises steeply at the
origin and is approximately straight (viscous range) but soon flattens out
1 hoc. cit. (second paper quoted there).
* Microplasticity in Crystals of Tin, Proc. Roy. Soc. London, ser. A, vol. 156, p. 427,
1936.

/T
68 THEORY OF FLOW AND FRACTURE OF SOLIDS

at large values of v. This function a = f(v) reminds us of the logarithmic


relation a = <n + o-0 In v/vo found by P. Ludwik in his experiments made
with polycrystalline tin wires, which were mentioned in Sec. 3-2. Chal
mers calls the initial viscous type of creep of a tin crystal "microcreep,"'
the normal extension at higher stresses "macrocreep." It is presumed
when single metallic crystals deform slowly under constant load that
their creep is due to slip. Since amorphous solids (glasses) not having a
crystalline structure creep at sufficiently high temperatures, creep in
amorphous solids must have a mechanism other than slip, which is
closely related to that of the mobility of viscous liquids. Amorphous
solid materials creep by virtue of their viscosity. But, to return to
metals, recrystallization in polycrystalline metals must have a part in
creep at sufficiently high temperatures. Alloys and metals with inclu
sions may have their own ways of creep. One structural change in
medium-carbon steel under prolonged loading has already been men
tioned (page 46), namely, the spheroidization of cementite accom
panying creep. From these remarks the conclusion may be drawn that
in the structure of metals under sustained loads at elevated temperatures
various changes have been observed: slip, deterioration of the grain
boundaries, recrystallization, spheroidization, and other metallurgical
changes. The ultimate common cause in all these phenomena accom
panying creep, however, is change of position of the atoms under their
thermal agitation.

A mechanism of change of position of atoms has been considered by H. Eyring in


his theory of diffusion and of the viscosity of liquids.1 It is based on physicochemical
concepts. The mechanism of viscosity during the laminar motion in a liquid is
viewed as a transport of the energy of thermal agitation through the change in posi
tion of atoms or molecules which is more freely possible in a fluid than in a regular
arrangement of these particles in a lattice in the solid state (and in distinction from
the transport of momentum of translatory motion of the atoms in a gas making up
the Maxwell mechanism of the viscosity of gases). Space for the jumps of the atoms
of a liquid during its laminar motion is provided by the concept of "holes" introduced
by Eyring, the atoms in a fluid being separated by more space than in their nearest
regular arrangement in the crystal lattices with their densest way of packing of
particles.
According to this theory of Eyring* the dependence of the rate of shear » = dy/dt
in laminar motion of a general fluid on the shear stress r and on the absolute temper
ature T is expressed by

v = ^
dt
= aTtr-fl inh
.
(yj
1 Loc. cit.

■ "The Theory of Rate Processes," p. 483. Sop also reference in the following
footnote and II. Eyring, "The Creep and Plastic Flow of Solid Materials," National
Academy of Sciences, autumn meeting. 1947.
MECHANISM OF FLOW IN GRAIN STRUCTURE 69

(ci.r-.fa const). If the shear stress t is very small in the substance the argument of
the term containing the hyperbolic sine function is small and cit/T may be substi
tuted for it. This gives a linear function between v and t

v = Ce-^rT

corresponding to a simple viscous fluid. If, on the contrary, ctr/T is large, i.e., for
substances which are able to carry high stresses, the hyperbolic sine term becomes
equal to J£ec"r/T and after taking logarithms

T = C, + C, In V

for the stress t a logarithmic speed relation is obtained. Such a function has been
established from experiments for the dependence of the yield stresses on the rates of
strain both at normal and at elevated temperatures in certain ranges of these variables
for polycrystalline metals. It is remarkable that this theory which originally was
developed for fluids would thus seem applicable to rate phenomena in solids.1
The theories of dislocations by Taylor and of rate processes by Eyring have been
combined for deriving an equation expressing the steady (minimum) rate of creep in
metals in terms of physical constants, the stress and temperature.* For pure metals
it is claimed that the rate of creep should decrease with increasing values of the
modulus of rigidity of the metal and that creep is caused by slip in single metal crystals.
1 Kauzmann, W., Flow of Solid Metals from the Standpoint of the Chemical Rate
Theory, Metals Technol., vol. 8, June, 1941.
* Nowick, A. S., and E. S. Machin, Nat. Advisory Comm. Aeronaut. Tech. Note,
April, 1946, 21 pp.
CHAPTER 8

THEORY OF THE TENSILE TEST1


8-1. Conventional Stress-Strain Curve.
We have already mentioned
that certain solids do not deform much under the action of tensile or
compression loads. Glass, natural rocks, concrete, and cast metals
behave in this manner at normal temperatures. They are commonly
designated as brittle materials. In this chapter we shall not discuss their
behavior but shall assume that we deal with ductile polycrystalline
metals. If a prismatic or cylindrical bar of ductile material is loaded
continuously and if we plot the stress
<To = P/A0 as a function of the unit

strain e = (I — l0)/l0 = Al/h (P =


load, Ao = area of original cross sec
tion, lo = original length, I = length
under the load P, Al = I — l0 = exten
sion of bar), we obtain a stress-strain
curve <to = /(e) of the material as
shown in the lower curve of Fig. 8-1.
When the load P or stress <r0 reaches
its maximum value, the unit exten
Fig. 8-1. Stress-strain and true stress
sions or strains e cease to remain uni
curve.
form. At this point the test bar
begins to show a local constriction, or neck, and is said to start to neck
down. After P or a0 have further decreased the test piece finally frac
tures at the narrowest portion of the neck.
Neglecting the elastic deformation and the very small permanent
change in volume during a tensile test, we may consider the volume of the
bar as constant. If we designate the actual area of the cross section of
the bar by A at the instant when the load is equal to P, we have
Al = Aolo. (8-1)
Since I = Z0(l + we obtain for
«),

.-lo
=
A

(8-2)
+
1

Concerning the tensile test and its applications to the determination of the
1

mechanical properties of the ductile metals see the valuable monograph by G. Sachs
and G. Fiek, "Der Zugversuch" (The Tensile Test), Akademische Verlagsgesell-
schaft m.b.H., Leipzig, 1926, 252 pp.
70
THEORY OF THE TENSILE TEST 71

and by dividing the load P by the area A the true tensile stress a in the bar
P P -40
= c„(l + t) , (8-3)
A0 A

which is shown in function of the strains t in the curve OB'C in the stress
diagram of Fig. 8-1. A continuation of the curve a = F{t) representing
the true stresses beyond point C" which corresponds to the maximum of
the load P or stress c0 cannot be plotted as long as the dimensions of the
minimum cross section of the test bar in the necked portion have not been
observed. From Eq. (8-3) follows that

<7o
=
1 +(
At the maximum of the load P or the stress c0 = P/A 0, dco/de = 0, the
bracket in Eq. (8-4) must vanish, and
do
— CO nun
(8-5)
dt 1 +*
We note from Eq. (8-5) that at the maximum of the stress c0 the slopo
da dt of the curve of the trite stress c = F(t) just becomes equal to the
ultimate stress coma, in the stress-strain curve. This is indicated also by
a construction shown in Fig. 8-2. The point C" on the curve of the true

1 c'
I

a
<r0

l/°
~
\

°6 MAX.

Ej,/'
'


0

'

+■- — e— +■
1

\r+
/

Fio. 8-2. Construction by Considfere.

stresses a corresponding to comn found by drawing the tangent to the


is

curve c = F(t) from point E.


a

situated on the axis at distance OE = — to the left from the


E
is

a

origin O.1

This construction mentioned in the book by A. RejtO, "Einige Prineipien der


1

is

theoretischen mechanischen Technologic der Metalle," p. 268, Vercin deutscher


Ingenieure, Verlag, Berlin, The construction attributed to Oonsiderc.
is

1927.
It may serve also in more complicated cases of deformation to locate the stress beyond
which necking or localized distortion starts.
a
72 THEORY OF FLOW AND FRACTURE OF SOLIDS

The maximum ordinate of the ordinary stress-strain diagram <r0 = f(t)


"
commonly known in the engineering literature as the "ultimate strength
has merely the meaning, as Ludvik emphasized a long time ago, that at
this point the uniform extension of a specimen ceases in a tensile test.
In the curve of the true stresses a = the stress a = DC (Fig. 8-2)
does not have a particular physical significance because no new condition
is defined by this stress in the material itself. The tangent construction

TRUE STRESS <T*% O


75,000


60,000

45fi00

30,000
, f hc=f
toccc eraa in r.

h
v'

ki 15,000
?!

O 5 10 15 20 £5 30
E+.UN1T ELONGATION IN PER CENT
Fig. 8-3. Tensile stress-strain diagram of wrought iron (annealed during 1 hr at 1000°C).

TRUE STRESS
45 mm 1
L -

id 3,5,
stre 55 STRtUN CUR VE <r-% _
13

UNIT STRAIN IN PER CENT


Fio. 8-4. Tensile test of copper (annealed during 1 hr at 8,0°C).

described in Fig. 8-2 however, perhaps significant from another view


is,

point: has served in the example of the tensile test to predict a limit for
it

the stability forming process in the sense that beyond the corre
a given
of

sponding stress or strain the deformation ceases to remain uniform. It


is

obvious that similar conditions may prevail under more complicated


systems of straining, e.g., when a tube expanded through internal pres
is

sure and simultaneously acting axial load, both increasing according to


a

some predetermined manner, and desired to predict the pressure or


it
is

load or both which will cause localized bulging.


The stress <to and the true stress a referred to the original and to the
actual cross section, respectively, are plotted against the unit extension
in tensile tests for wrought iron in Fig. 8-3, for copper in Fig. 8-4, and for
aluminum in Fig. 8-5.
THEORY OF THE TENSILE TEST 73

Attempts to define and construct the curve of the true stresses beyond
the beginning of necking in tensile tests have frequently been made. The
use of the unit strain « computed from the extension Al of a, finite length l0
is not suitable for this purpose since it measures only the average value
of the strains after necking started. Koerber1 and other investigators
pointed out that a mean strain « could be calculated in the contracted
portion of the test bar if the changes of the diameters were observed. It
was suggested that the reduction in area in the minimum section of the bar

TRUE STRESS <r • %

4-
STRESS STRAIN CURVE <r„ - '
I I I °_L
5 10 1520 25 JO 35
E-+-UNIT ELONGATION IN PER CENT
Fig. 8-5. Tensile teat of aluminum (annealed during 1 hr at 350°C).

q = (Ao — A)/Ao be taken as the variable for measuring the deformation.


Since A0/A = 1 + t, the mean strain t in the minimum section would
be found

and thus the stress-strain curve would be extended beyond the necking
point to fracture. A more satisfactory method for constructing this
portion of the stress-strain curve will be considered in the following sec
tion, however.
8-2. The Natural Stress-Strain Curve. By definition the strain
t = (I

lo)/lo- This is equivalent to

(8-6)
<o Jh
Instead of defining the increments of strain by de = dl/l0 P. Ludwik and
A. Leon2 suggested in 1909 that a new strain e be defined by the increments
dt = dl/l or by
/.7

= = = ln(1+«)- (8-7)

This has the great advantage that the condition of the incompressibil-
ity of metals (or of other incompressible substances such as rubber) can
be expressed in a simple manner even though the strains may become

Korber F., MM. Kaiser-Wilhelm-Inst. vol. Part


1

Eisenforsch. Dxisseldorf,
2,
3,

1924.
Ludwik, P., "Elemente der technologischen Mechanik," Verlag Julius Springer,
*

Berlin, 1909.
74 THEORY OF FLOW AND FRACTURE OF SOLIDS

large. Suppose that a prismatic brick-shaped element of material is


deformed so that the right angles do not change; that ao,bo,co are the
lengths of the edges in the undistorted and a,b,c in the deformed state.
By definition ea = (a — a0)/ao, ft,, ec are the strains along these three
perpendicular directions. The constancy of volume
abc = aaboCo (8-8)
is also expressed by
+ e.)(l + + =

«e)
(1 «»)(1 (8-9)

1
or after taking natural logarithms by
+ In In + =

f„)
+
In

+
(8-10)

(1

(1

0
«.)
(1
««)

.
But this the same as
is

+ + = (8-11)

ia

0
the new definitions are used for strain. If the strains ta,tb,tc were
if

infinitesimal, the same condition

+ + tc =

0
«o «6

would have expressed incompressibility after


neglecting the higher powers of the strains.
Thus we note that this simple form of the
B
A

Flo. 8-6. Extensometer re- j... - ., .1. « . ~ ..


condition ofe incompressibility for infinites

,
cording the natural strains.
imal strains preserved also for finite de
is

a
formation provided that the strains are defined according to Eq. (8-7).
The author proposed to call the quantities ia,h,ic the natural strains in
distinction from the conventional strains «a,«6,ee in a strained material
element.1 The extensometer sketched in Fig. 8-6 will record on its
circular scale D having uniform divisions the natural strain in material a
stretching between the point and the contact point of the roll
B
A

C
in the direction AB. An extensometer built after Bauschinger's original
suggestion at the Westinghouse Laboratories for measuring very small
strains and based on this principle, but which suitable for recording
is

large uniform strains, reproduced in the photograph in Fig. 8-7.


is

Another advantage of using the natural strains instead of the con


e

ventional ones that the former ones are the variables which can serve
is

to define the true plastic rates with which infinitesimal elements in


a

material stretch or contract. Suppose that cylindrical bar stretched


is is
a

by moving one end with the velocity u while the other end fixed and
that the length between these points at the time and at = 0.
is

t
/

Then u = dl/dt. The rate with which the conventional strain


-
U
I

J. Applied Phys., 1937, p. 205.


1
THEORY OF THE TENSILE TEST 75

changes with time t

_ dt _ 1 dl _ u
dt /o d< to

is proportional to the velocity u between the heads of the bar. The true
rate of flow in the bar with which a unit length measured on the bar at the
time t changes, however, is v = di'dt:
1 dl u 1 dt
(8-12)
I dt 7 1 +€ dl 1 +

This rate of stretching we shall call the natural rale of strain.


v If the
velocity between the heads of the bar u — const, v must decrease in pro
portion as the length I of the bar increases. The mean natural rate of

Fig. 8-7. Extensometer measuring natural strains.

contraction or of necking in the lateral direction in the minimum cross


section of the bar is one-half of its natural rate of stretching [as can be
seen from Eq. (8-11) by assuming h = «c]. The natural rate of strain
in the axial direction in the neck of a bar is therefore equal to
. =
di IdD
V
dt ~lD~dt'
where D is the minimum diameter in the neck. '

1 If instead of the conventional reduction in area q = (A0 —


A)/A0 one introduces
A0
a natural reduction in area denned by q = — / —r- — In A*/ A = 1 +

one sees that this is equivalent to plot q = i for the abscissas in the stress-strain curve
or to make use of the natural strains I.
76 THEORY OF FLOW AND FRACTURE OF SOLIDS

The natural tensile stress-strain curve of a material, in which the true


stresses a are represented as a function of the natural strains I, is obtained
by retracing the true stress-strain curve a = of the conventional
diagram on semilogarithmic paper in which the abscissas are proportional
to
I = log. (1 + e)
= ln(l + e)
= 2.303 log,0 (1 + «) .

This is illustrated by two examples. The pair of stress-strain curves


shown in Figs. 8-8 and 8-9 may refer to a steel-like material. The second

/Ox/04 LB./ IN* IQxIO4 LB. /IN*


STR ESS ^
TRUE
-SSO I 1
fit
JfU;

ZON
< v.s TRt UN f-
> 0.2 0.4 0.6 0.8 1

CONV. STRAIN £-**


O 0.2 0.4 0.6
NAT. STRAIN f_
Fig. 8-8. Stress-strain and true stress curve. Fio. 8-9. Natural stress-strain curve.

set of curves (Figs. 8-10 and 8-11) should show the stress-strain behavior
of an ideally elastic and incompressible substance. The elasticity law
for this substance was assumed to be

^n
a = Ei , <r0
= E y-t~
~ = Eie-% , (Poisson's ratio v =

i.e., the true stresses a were assumed to increase proportionally with the
natural strains e. (In analogy to Hooke's law for infinitesimal strains
the proportionality factor was designated by E. E would be the modulus
of elasticity if the strains were infinitesimal and E should remain a con
stant if the strains become finite.) We may note from the Figs. 8-10 and
8-11 that the load for such an ideally elastic incompressible material
should become a maximum at the natural strain e = 1 or the conven
tional strain t = e — 1 = 1.718, also that specimens of such a material
could be stretched to two to three times their original length under a
practically constant load <r0mn, = Ee~l = 0.3G8B. The true stress is
a = E when an — <r0m.i.
The natural stress-strain curve for ductile metals may sometimes be
approximated in its first steep portion by assuming that the true stress <r
THEORY OF THE TENSILE TEST 77

is a power function of the natural strains i or that

a = <ri(e)n (n < 1)

where <r\ and n are material constants. The stress <ro referred to the
original area of cross section A0 is

<ro
=
1 + e

The condition of instability, under which necking starts, is expressed by

1.5

%
i.o
Fig. 8-10.
\
MAX.
0.5
I
(To X

Y 2
CONV. STRAINS

f
-%
'E

f
MAX.
Fio. 8-11.

/ - 2
NAT. STRAINS
Figs. 8-10 and 8-11. Stress-strain curves for ideal elastic incompressible material.

making this stress a0 an analytic maximum or dao/di = 0. This leads


to the condition i = n. At this natural strain I a tensile bar would cease
to stretch uniformly.1

W. T. Lankford and E. Saibel (Metals Technol., August, 1947) made use of such
1

power functions under biaxial conditions of loading in plates and in hollow thin-
walled cylinders under combined axial loading and internal fluid pressure for predict
ing the combination of the two principal stresses under which the uniform extension
ceases in such bodies. They also predicted similarly an approximate condition under
which a thin circular diaphragm of a metal bulged laterally by a fluid pressure should
- to stretch uniformly.
78 THEORY OF FLOW AND FRACTURE OF SOLIDS

MacGregor1 and Davidenkov2 made extensive use of the true stress-


natural-strain curves in their comparative studies on ductile metals.
Both found that these curves beyond the point at which the necking
starts are nearly straight lines. This led both investigators to suggest
further simplifications in the method of making tension tests. Mac
Gregor3 proposed "a two-load method" for determining the stress-strain
curve, whereby the diameters at various sections along gently tapered
round bars are measured before and after test and only the maximum and
the fracture load are recorded. This can be done without interrupting
the test because those portions of the bar which carried stresses which
were smaller than the true stress corresponding to the maximum load
cease to deform after the load starts to drop. For the remaining portion
of the diagram a straight line is substituted in accordance with other
tests. This method simplifies recording the extension in tests at elevated
temperatures or in impact tests.
For obtaining firsthand information concerning the position of the
straight-line portion of the true stress-natural-strain diagram for a
ductile metal Davidenkov suggested making a Martens scratch and a
Brinell ball hardness test on a piece of the metal, eliminating the necessity
of running a tension test at all including the making of a tensile test
specimen. Available data correlating the increase in hardness with cold
work permitted him to make conclusions concerning the stress-strain
curve.
Through cleverly varying the velocity between the heads of a testing
machine during a tensile test MacGregor and Fischer4 were able to realize
tests run under a constant mean natural rate of strain (v = const) in
the minimum section of a bar and to study the effect of this rate on the
stress-strain curve for a 1020 and 1045 steel and brass.
The work done by the load P for stretching permanently a prismatic or
cylindrical bar in tension is equal to

W = fP dl (8-13)

1 MacGregor, C. W., The Tensile Test, Proc. ASTM, vol. 40, p. 508, 1940. This
paper, as well as several related papers, which were presented in a symposium on
the tensile test arranged by the ASTM, contains also complete bibliographies on this
subject; the other papers include effects on the speed of stretching for metals, etc.
1 In a number of papers published in Russian journals and in several monographs.
* J.Applied Mechanics, 1940. Cf. also two papers by A. V. deForest, C. W.
MacGregor, and A. R. Anderson and by MacGregor and L. E. Welch on applica
tions of the two-load method to rapid and to high-temperature tests in Metals Technol.,
1941 and 1942.
4 Tensile Tests at Constant True Strain Rates, ./. Applied Mechanics, 1945.
THEORY OF THE TENSILE TEST 79

where U and I denote the original length and the length under the load P
of the bar. After substituting here for P = <r0A0 and for I = lo(l + e),
dl = h dt in accordance with Sec. 8-1 the total work of deformation is
expressed also by

W = A„l0 (8-14)
fo'<r0de.

Since A o/o is the volume of the bar, we see that the integral in the preced
ing equation must express the work of deformation per unit of volume.
Denoting the latter by w, we see that

w = (8-15)
f0'<rode

is found by computing the area under the conventional stress-strain curve


at,= /(*) of the material tested in tension.
Instead of using the load stress a0, we may introduce in Eq. (8-15)
the true stress <r by substituting according to Eq. (8-3) for <70 = o-/(l + *)
and may make use of the natural strains i = In (1 + «). Since

the work per unit of volume is also expressed by

w = L'cdi (8-16)

or by the area under the true stress-strain curve a = F(i) plotted above the
natural strains i.
8-3. Stress Distribution in Neck. When the test piece begins to neck
after the maximum load is reached, a rotationally symmetric distribution
of the stresses gradually develops in the constricted portion of the bar.
This complicates considerably the precise evaluation of the stresses in
ductile materials at or near fracture. A useful approximate calculation
of the stress distribution in the neck of a test bar was given by E. Siebel.1
Since the principal stresses which act in the longitudinal direction near
the surface of the bar cannot remain parallel to the axis of the specimen,
there results a radial pull directed outward in the minimum section in the
neck and a nonuniform distribution of the stresses in the axial direction
in it. These inclined longitudinal stresses create in round bars in the
center region of the minimum section a state of triaxial tensile stresses
consisting of axial, radial, and circumferential stress components which
depend on the distances of a point in the minimum section from the axis
of the bar, the radial stresses vanishing at the circumference.
1 Dtisseldorf, Slahl
Werkstoffausschuss Ber. 71, Verein deutscher Eiscnhlittcnleutc,
u. Eisen, 1925.
80 THEORY OF FLOW AND FRACTURE OF SOLIDS

By using Armco iron with very distinct grain boundaries N. N. Davi-


denkov and N. I. Spiridonova2 were able to study the distribution of the
natural strains in the necked portion of such bars by etching longitudinal
sections in them. From these careful studies they determined an approx
imate distribution of the stresses. These investigations led them to the
conclusion that the natural strains in the minimum section in the radial
and in the tangential directions must be equal and constant through the
entire cross section. This then justifies the assumption on which to
base the further computation of the distribution of the actual stresses
near the necked portion of a tensile bar. These experiments make it
possible to use a mean natural strain i (mentioned above in the text) as
an independent variable in the stress-
strain curves also beyond the start of
AXIAL STRESS
necking.
The last mentioned two authors sug
gest a correction term which must be
included in the formula expressing the
value of the true free tensile stress <r„ in
the minimum section, which must be
plotted as the ordinate corresponding to
the strain e in the tensile diagrams. <ra

is less than the average true stress in


this section. If P is a load at which a
Fio. 8-12. Stress distribution in constricted portion formed and a is the
neck of tensile bar.
radius of its minimum section, the mean
stress in it is equal to am = P/ira*. Furthermore, if R is the radius of
curvature of the profile of the neck (Fig. 8-12), the true free tensile stress
<ra (the tensile stress at the surface
r = a of the bar in the neck) is to be
taken equal to

1 + (a/4R)
For r = a according to these authors, since everywhere in the minimum

section it = I, = —ia/2 the circumferential stress at must be equal to the


radial stress ar, which is equal to zero at the surface of the bar.
At the center of the minimum section the axial stress is equal to
R + 0.50a
= On
R + 0.25a
1 "Analysis of Tensile Stress in the Neck of an Elongated Test Specimen," paper

published by Scientific Research Institute of Aviation Materials of Soviet Govern


ment, Moscow, 1945 (English translation by Westinghouse Research Laboratories),
Proc. ASTM, 1946.
THEORY OF THE TENSILE TEST 81

This is the value of the maximum stress in a tensile bar in the smallest
section of the neck at which the facture must start if P is the fracture
load.1
P. Ludwik2 was probably one of the first who noticed that the fracture
in round bars of a ductile metal under a tensile force starts at the center
of the minimum section in the neck. By taking a photograph of the
necked portion of an aluminum bar by means of X rays just before the
bar broke in two pieces, C. W. MacGregor3 (Fig. 8-13) verified this.
That the fracture in round bars of ductile metals in ten
sion tests starts with a crack at the center of the min
imum cross section is proved by machining a longitudi
nal section through the axis of the bar after the broken
ends are fitted together again. A distinctly visible gap
appears in the central portion of the bar between
the two broken halves (Fig. 8-14, copper) showing that
the outer portions in the "cup and cone surface" were
still stretching after the central portion already had
separated.
Further examples of the widely varying profiles of
round bars broken in tension are illustrated in Figs. 8-15
to 8-17. Figure 8-15 shows the broken halves of a large FlO. 8-13. First
medium-carbon steel bar which had a cylindrical length crack in round
bar of aluminum
of 15 in. and an original diameter of 3 in., Fig. 8-16 tested in tension.
shows two small aluminum bars which were pulled at (X-ray picture by
MacGregor.)
600°C and necked down practically to a point, and Fig.
8-17 shows the necked portions of three stainless-steel
bars which had an original diameter of 3^ in., a cylindrical gauge length
of 8 in., and which were tested at 454°C (850°F) by pulling under con
stant rates of strain. The rates are shown in the legend to Fig. 8-17
in inches per inch per hour. While the bar represented in the middle
of Fig. 8-17 broke at an elongation of 18 per cent after 276 min,
that on the right broke at 10 per cent after 7.5 min. The slowest test
had a duration of many days. It is interesting to note that in the fastest
test (right bar) the necked portion was much shorter than in the slowest
test (left bar).

1 A set of analogous formulas has been proposed by P. Bridoman, Trans. ASM,


vol. 32, p. 553, 1944.
« Z. Ver. deut. Ing., vol. 71, See also Ludwik, P., "Bruchgefahr und
p. 1532, 1927.
Materialpriifung," Schweiz. Verband Materialpriifung. Tech., Ber., No. 13, Novem
ber 1928, Fig. 3, p. 29, Zurich, 38 pp.
* Relations between Stress and Reduction in Area for Tensile Tests of Metals,
AIME Tech. Pub. 805, New York meeting. 1937.
82 THEORY OF FLOW AND FRACTURE OF SOLIDS

Fig. 8-14. Copper bar tested in tension. Fig. 8-15. Broken round bar of a medium
(The bar cracked first in the center of the (0.23 per cent) carbon steel. Original diam
minimum section. Note subsequent eter 3 in. [Acknowledgments are due to
formation of a gap in the central portion J. Aronofsky (Pittsburgh) and to the David
of the tensile specimen. Longitudinal Taylor Model Basin, Navy Department,
section through axis of bar. The test Washington, D.C. for permission to reproduce
was interrupted just before the round this figure. See also two papers on the si«e
bar broke.) [Acknowledgments are due effects in round and flat bars tested in tension
to J. Miklowitz for making the test and by J. Miklowiti (J. Applied Mechanics, 1947,
to R. L. Anderson for etching the section. 1948).]
The lines appearing on the polished sur
face (oxide inclusions) disclose the flow
after necking.]

W. von Mollendorff1 claimed that the profiles of round bars in their


constricted portion could be represented by two surfaces of revolution:

r = a(l - «****>'•)
< + as , both upper signs\
(—c^x
— < x ^ c, both lower signs
oo
/
where r is the radius and x the coordinate in axial direction; r = a when
x = + oo, and the constant c depends on the material. These two con
1 Mill. M aterialyrufungsamt Berlin-Dahlem, vol. 41, Nos. 5, 6, pp. 51, 60, 1923.
THEORY OF THE TENSILE TEST 83

gruent surfaces have to be rounded off by a short hyperboloid of revolu


tion in the middle portion of the bar (Fig. 8-18). The more ductile a

is,
metal the larger the constant becomes (2c measures the distance

c
between the points O\ and 02 in Fig. 8-18). The formula checked well
with the profiles of the necks of bars of soft steel and copper, semihard

a
steel and lead.

The distribution of the stresses in the


minimum cross section of a round bar of a
hot-rolled 0.25% steel at fracture has
C

been determined by E. R. Parker, Harmer


E. Davis, and A. E. Flanigan1 by a method
in which the residual stresses in this sec
tion were observed in tension tests.

A
series of round bars were unloaded after
the necking had partly started in them,
and metal was removed from the interior
of the specimens by successively drilling
larger and larger holes in the axial direction
in them while the elastic changes of the
strains during the redistribution of the
residual stresses were observed by means
of wire resistance strain gauges which were
attached to the bar on the outer surface
of the neck at the minimum diameter.
This method for a determination of the
residual stresses in pieces which have been
plastically deformed in nonuniform man
a

ner will be described in Vol. 2. The


Fio. 8-16. Aluminum bars pulled at
authors assumed, as most investigators do, 600°C. (A/. J. Manjoine and author, in
that during the unloading process stress J. Applied Mechanics, June, 1941.)
and strain are related to each other
through the equations which are valid in an elastic body. In these computations
they used certain formulas which had been derived by H. Neuber for elastic
bodies having the shape of hyperboloid of revolution.1 In this manner they were
a

able to compute the original distribution of the stresses in the condition of the bar
before the load was relieved while was yielding in the necked portion. The distri
it

bution of the radial, tangential, and axial stresses ar, at, and a, in the minimum cross
section of the neck in function of the radial distance of a point from the axis of the
r

bar had a great resemblance to the one which was computed by Davidenkov and
Spiridonova (see Fig. 8-12), although certain deviations appeared in the correspond
ing curves for o>, at, and aa which were due to the unavoidable experimental errors of
this method. The authors also described valuable observations concerning the
cleavage and shear fractures in the grain structure of the bars to which further refer
ence will be made in Chap 15. They found that the fractures in the crystal grains in
the tension specimens of mild steel tested at normal temperature were both in the

Study of the Tension Test, Proc. ASTM, vol. 46, pp. 1159-1174, 1946.
1
A
1

hoc. cit.
84 THEORY OF FLOW AND FRACTURE OF SOLIDS

inner and in the outer portions of the cup and cone fractures of the shear type. In
specimens which were broken at liquid air temperature the crystal grains showed the
cleavage-type fracture, while the ferrite grains at intermediate temperatures broke
by shear as well as by cleavage fractures.

It is believed that a mechanical theory of the necking of ductile metals

which is based on the concept of pure strain hardening will not predict
one other important group of observations related to necking, namely,
the difference in the shapes of the profiles of the necked portions of tensile
specimens which has been observed when bars are tested under different
rates of stretching at an elevated temperature. We have called attention
THEORY OF THE TENSILE TEST 85

to two examples in Figs. 8-16 and 8-17. ' Two Russian investigators,
A. A. Ilyushin2 and A. J. Ishlinsky,3 have recently worked out a mathe
matical theory of necking assuming certain idealized conditions, namely, a
material which has a definite yield point and which, after this has been
reached, deforms like a viscous material (in which stress is proportional
to the permanent rate of strain). They found the interesting result that
bars of such an ideal material having a profile with shallow waves may or
may not neck farther down in the narrowest sections. This depends on
the rates with which the bars are stretched. If the rate is below a certain
rate, the shallow waves are wiped out through further stretching. If the
rate is above the critical value, the waves will become deeper and deeper
and necking will proceed in the bars.
It is of interest to note that the radial and the tangential stresses aT
and <r< in function of the radial distances r which were determined by
Parker, Davis, and Flanigan at the instant when the fracture starts in
the minimum cross section of the bar were represented by similar curves
within the experimental errors so that their tests verified the assumption
made by Davidenkov and Spiridonova that a, = at and that furthermore
tr = d = const in the minimum cross section. Such a special mode in
the flow discloses a strikingly simple rule according to which the dis
tortion in the necked portion of a tensile specimen proceeds and must
have a deeper mechanical cause. Suppose that the condition of plas
ticity is expressed by an equation:

+ — a ay
+ — = = const
(ov (?«)* (<r» (o-a <rr)s 2o-02

or by "a surface of yielding" in an advanced stage of strain hardening


of a ductile metal approaching the last stages of stretching just before
the fracture may occur. * o-0 in this equation is a constant. Assuming that
<t„, o>, and at (the axial, radial, and tangential stress at the distance r from

the axis in the minimum section) are plotted as rectangular coordinates


(see accompanying Fig. 8-19) this is the equation of a circular cylinder
whose axis makes equal angles with the <?a, oT, and <r, directions. On the
1 A few examples in which the contour lines of round bars were computed during
subsequent stages of the necking process have been discussed under simple assump
tions as to the speed laws in a paper by M. J. Manjoine and the author in Applied J.
Mechanics, June, 1941.
* Ilyushin, A. A., Deformation of Plastico-viscous Materials, Scientific Notes of
the State University in Moscow. U.S.S.R., Issue 39, 1940.
* Ishlinsky, A. J., Stability of the Plastico-viscous Flow of a Round Bar, Russ. J.
Applied Math. Mechanics, vol. 7, No. 2, pp. 109-130, 1943, Moscow; also Sokoi.ovsky,
W. W., The Theory of Plasticity — An Outline of Work Done in Russia, /. Applied
Mechanics, vol. 13, No. 1, p. Al, 1946.
4 See Chap. 15, p. 210.

"
Kli THEORY OF FLOW AND FRACTURE OF SOLIDS

right of the cylindrical surface of yielding is shown a portion of a necked


bar in two views (Figs. 8-20, 8-21). The testing engineer wishes to
express the distribution of the axial stresses <ra within the minimum sec
tion (as indicated in the middle figure through the curve for <ra) of the
neck.
The cylindrical surface of yielding intersects the coordinate planes in
two ellipses shown in Fig. 8-19. To point A of the bar (at the surface

(AXIAL
STRESSh

(RAD.
STRESS)

Fig. 8-19. Fio. 8-20. Fiq. 8-21.


Fia. 8-19. Surface of yielding in advanced stage of plastic straining of metal.
Fio. 8-20. Axial stresses in minimum cross section. Fio. 8-21. Minimum cross section
of bar.

of the neck) must correspond a point A situated on one of these ellipses,


namely, in the plane section of the surface of yielding in which <jt = 0
(since the radial stress <rr must vanish in point A on the surface of the
tension specimen). At point C, the center of the minimum section (or in
the axis of the tension specimen), a, equals crt. The corresponding point
C (representing the state of stress) on the surface of yielding must evi
dently be situated in the plane bisecting the right angle of the o> and <r,
axes (since in this plane o-r = ct). Thus we see that all states of stress in
the minimum section of the tension specimens will be represented by a
curve which connects the points A and C on the cylinder. This curve
has been traced in the accompanying Fig. 8-19 on the cylindrical surface
of yielding. The ordinates o-„ of this latter curve (such as CC) then rep
resent the axial stresses <ra of the bar in its minimum section just before
fracture occurs.
It is now easy to see that the integral of all these ordinates,
THEORY OF THE TENSILE TEST 87

-*£ Jo
aar dr

which is the load at fracture becomes a minimum if the curve CA is made


to coincide with the steepest straight line CB which can be traced on the
cylinder, namely, the generatrix passing through C. All other curved
lines traced on the cylinder starting from point C and joining a point A
in the aa,at plane have longer ordinates aa then the straight generatrix
CB. But along this latter
o>
= at .

Thus we may state that according to the theory of flow of a perfectly


plastic substance a ductile metal adjusts itself to flow under tension in
such a manner that in the minimum cross section in the neck everywhere
the radial stress becomes equal to the circumferential stress because
under these symmetrical states of stress the bar can flow under the mini
mum value of the tensile force, which otherwise could make it also flow if
aT were not equal to the at.
In a book by F. K. Th. van Iterson1 attention has been called to this
symmetrical type of flow in which two principal stresses (a„at above)
tend to become equal to each other. When one head of a tension bar is
moved relative to the other head with a given constant velocity, since
during the last stages of the tensile test most of the distortion is localized
near the minimum cross section of the specimen, one sees that the sym
metrical mode of flow in which ar = at requires the least amount of
mechanical work to be done to deform the bar, or, since in this region
the strains ta uniformly increase, that the bar will deform plastically in
the manner which will make the axial force P to a minimum.
8-4. Mechanical Similarity. To ensure mechanical similarity in the process of
plastic deformation in two geometrically similar bodies which are loaded only by
forces acting on their surfaces (inertia or body forces such as the weight are excluded)
the stresses in corresponding points in the interior of the bodies and at corresponding
times must be the same. True mechanical similarity is ensured if the trajectories of
principal stress in the bodies are families of geometrically similar curves in space.
If, for example, two geometrically similar cylindrical or prismatical specimens are
stretched by tensile forces and if it is known that the tensile stress-strain curve depends
also appreciably on the rates at which a test bar is permanently elongated, to ensure
equal plastic strain rates in corresponding points, the relative velocities with which
the heads of the bars are displaced must vary in the same proportion as their dimen
sions. A large specimen having the same ratio of length to diameter in the gauge
length as a small specimen must be stretched with a relative velocity between its
heads which is as many times larger than the rate between the heads of the small
specimen as the diameter or length is larger than the corresponding dimensions of the
small specimen.
'"Plasticity in Engineering," Chap. 13, Blackie & Son, Ltd., Glasgow, 1947,
174 pp.
88 THEORY OF FLOW AND FRACTURE OF SOLIDS

It is well known that the elongation at fracture depends on the ratio of the length
to the diameter of a cylindrical bar tested in tension. If this ratio is larger than a
certain value, the manner in which the constriction forms in the bar after the ultimate
load has been reached is independent of the length. This is no longer true for shorter
bars. To ensure an undisturbed mode of necking the length must be at least ten
times the diameter.
A cylindrical and a fiat test bar behave alike mechanically only in the region of
uniform extensions; after the occurrence of necking the distribution of the stresses in
the constricted portions of both test pieces will be entirely different. In the German
standards (D.I.N. 1605) a rule for the determination of the gauge length U of a flat
bar having a cross section A used to be

U = 11.3 vT.
This length was obtained by prescribing that the ratio of l0' to the length of the
standard cylindrical bar should be as A to rd^/4 after taking as length of the cylin
drical bar W = 10rfo. As stated above, beyond the beginning of necking no strict
mechanical similarity of the process of stretching exists between a round and a flat
bar. We shall see later that the shape of the constricted portion in the neck of a
tensile specimen depends on a number of conditions. Tensile test bars tested under
different rates of strain neck down in different manners. This is particularly the case
at elevated temperatures.
CHAPTER 9

STRESS

Let us consider a body in equilibrium under the action of external


forces. Taking a cross section of this body and investigating the equi
librium of the forces acting on each part thereof, we see that, in general,
forces from one part of the body will be transmitted across the section
to the other part. The ratio of the force acting on a small element of
the cross section to the area of this element approaches a limiting value, if
this area is taken smaller and smaller. This limit is called the "unit
stress" at the point considered. This stress may be resolved into two
components, one perpendicular to the cross section and the other in the
cross section. The former is called a normal stress; the latter a shearing
stress. Normal stresses are designated by the letter a, shearing stresses
by the letter t; normal and shearing stresses in general will be distin
guished by subscripts attached to the letters a and r. Normal stresses
are taken positive or negative depending on whether they are tension or
compression while the sign of the shearing stress components will be
arbitrarily fixed for each case. In the bodies it will in general be neces
sary also to consider continuously distributed body forces due to inertia
or weight of the mass.
The distribution of the internal stresses around any given point in the
interior of a body is known, if the components normal and in the tangen
tial direction at this point are known for any arbitrary position of the
cross section. We may locate the point by means of its coordinates
taken with respect to a right-handed rectangular system of axes x,y,z.
The position of the cross section under consideration is determined by the
angles which the external normal to this section makes with the positive
directions of the x,y,z axes.
We shall first consider homogeneous states of stress by assuming that
the stress within the region to be investigated has the same magnitude
and direction in every point of a cross section. Assume a small element
in the body cut out by planes parallel to the planes of the coordinate axes
and having the edges dx,dy,dz. On the face of this element parallel, for
example, to the y,z plane, whose external normal points in the direction
of the positive x axis, a normal stress a, will act. The shearing stress in
this plane may be resolved into its two components and t«, parallel
89
90 THEORY OF FLOW AND FRACTURE OF SOLIDS

to the y and z axes, In this method of attaching subscripts


respectively.
to the letters designating components of stress, it will be noted that the
single subscript of a normal stress and the
first subscript of a shearing stress correspond
to the direction of the external normal to the
section, while the second subscript of t indi
cates the direction in which the component is

/
i f to be taken. We thus have the following
dz

A,
~y» / nine stress components parallel to the three
coordinate axes
f, T„
dy Ty, Oy Ty,
Via. 9-1. Components of
stress.
Their positive directions are shown in Fig. 9-1
for the three faces of the element the external normals of which point in
the directions of the positive x,y,z axes. From the condition of equilib
rium with respect to moments of the forces transmitted through the six
faces of the element dx dy dz, the
six components of the shearing
stresses must satisfy the equalities:

If
we consider a tetrahedron cut
out from the body by three planes
parallel to the planes xy, yz, and zz
and a fourth plane oblique to the
coordinate planes, from the condi
tion of equilibrium of the forces
acting on the four faces of this tet
rahedron the components of the
stresses in the oblique plane may
be determined. If we choose the
area of the oblique section equal to Fio. 9-2. Stress components acting on
tetrahedron.
unity, the areas of the sides of the
tetrahedron parallel to the coordinate planes, yz, zx, and xy are equal to
the direction cosines a,, a,, and a„ respectively, of the normal n to this
oblique cross section. After denoting by s,, s„, and s, the components of
the resultant stress s of this plane, the equilibrium of the forces acting
on the tetrahedron1 may be expressed by the following three equations:
1 In
the sketch (Fig. 9-2) only the components of stress acting in the oblique section
and in one of the coordinate planes (the x,z plane) were traced (those acting in the
x,y and y,z plane were not shown).
STRESS 91

sx = ajax + ryxay + Tzza, ,

8y — Tzydz + OyOy + M, , \ (9-1)


s, = rztaz + Tvra„ + a2a* .

We now resolve the stress s acting on the oblique face into its normal
and tangential components a and t (omitting to affix subscripts to these
stresses). Since the normal stress <r is equal to the sum of the projections
of the stresses sz, 8„, and sx, on the normal n to the oblique plane, we have

a = SjOx + SyOy + SjOj (9-2)

and the shearing stress r is given by

T2 = s- - <r2 = sx2 + sy2 + s,2 - <r! . (9-3)

Substituting in the first of these equations the expressions for sx, sy, sz
given by (9-1), we obtain for the normal stress:

a = (TjOz2 + fuCLy2 + afiz- + 2Txyaxay + 2Tyxaya, + 2i-IIaiax . (9-4)

The six components of stress ax,<Ty, ...


in Eq. (9-4) appear as the
coefficients of an expression of the second degree in the direction cosines
ax,Oy,az. By a suitable transformation of the direction of the coordinate
axes x,y,z through a rotation around the origin one may find a system
x',y',z' for which the three new components of stress r^y, iw, and Tjv all
vanish.1 Assuming that this system has been chosen, Eq. (9-4) for the
normal stress a takes the simpler form :

ff = ax2ai + ay2ai + a^Vg . (9-5)

If from a fixed
1 point 0, we lay off a distance OQ = r parallel to the normal (o„
Of, a,) and choose the length r such that its square is inversely proportional to the
normal stress: a = c/r1, after a suitable choice for the sign of the constant c, since the
components of the length r are x = a,r, y = ayr, z = a,r, Eq. (9-4) may be rewritten
as follows:
<rxx% + (Tuy* + <Trt,%+ 2rxyxy + 2ry,yz + 2r,xzx = c. (9-4a)

This is the equation of a surface of the second degree in the coordinates x,y,z. From
it we may see how the components of a state of homogeneous stress are changed if
the direction of the axes x,y,z is changed into x',y',z'. In the case of a rotation of the
coordinate axes the six stress components are changed in the same way as the con
stants in the equation of a surface of the second degree are changed when the coordi
nate axes are rotated. Since such a surface in general (if it does not degenerate into
a sphere, etc.) has three mutually perpendicular principal axes, for which the terms
containing the products xy, yz, and zx vanish in the equation of the surface, we must
conclude that, for a state of arbitrary homogeneous stress, there are three mutually
perpendicular directions in the body — the principal axes of stress — corresponding to
which Eq. (9-4) takes the simpler form of Eq. (9-5).
From Eqs. (9-1 ) one sees also that, if an oblique plane should carry only a normal
92 THEORY OF FLOW AND FRACTURE OF SOLIDS

In
the cross sections perpendicular to these three directions, the normal
stress a reaches a maximum or a minimum value and the shearing stresses
vanish. These three stresses <n, m, and o 3 are called the principal stresses
and their directions the principal axes of stress.
If we choose the principal axes of the state of stress as the coordinate
axes, Eqs. (9-1) simplify to the following form:

<j\ax Sy = (Tilly S, = C3CI1 (9-6)

If we substitute the values of the direction cosines az,Oy,a. computed


from the preceding three equations in

1 (9-6a)

we obtain for the components sz,sy,s, of the resultant stress s:


2 .. - r- 2
8
- 2 ' _ 2 '
„_2
A (9-66)
<Tl a2" 03-

or in the variables sx,su,sI the equation of an ellipsoid. This is called the


stress ellipsoid. It represents the loci of the end points of the resultant

stress o, but no shearing stress, the components »*, s„, s, must be sx = aa„ sv = aa„
s, = on,. Thus for a principal plane of stress Eqs. (9-1) become a system of three
linear homogeneous equations in a., a„, a, the determinant of which must vanish, or

= 0. (9-46)

This is a cubic equation in the unknown c, which has three real roots: 01, a~, <ri (see
Frank, P., and R. v. Mises, "Die
Differential und Integral-gleichungen der Mechanik
und Physik," Vol. 2, p. 605, Friedrich Vieweg & Sohn, Braunschweig, 1927), the three
principal stresses. Since the latter cannot depend on the coordinate system, one sees
at once' that the coefficients of the cubic equation determining the three principal
stresses must be "invariants" of the state of stress. After evaluating the determi
nant Eq. (9-4b), one finds that the coefficients of the terms containing a1, a and the
constant term in the cubic equation or the three invariants are equal to

Ty* T,,
"z"y + O yO, + tt&x (Tx„S + TV> + T„») , a,j T,v (9-4r)
Ty* 0,

If the directions of the X,y,i axes coincide with the principal directions of stress
the shearing stresses vanish in these expressions. When expressed by means of the
principal stresses <n, <r*, at these same three invariants take the form:

'l + Ot + "l , fflffl + 0-2^1 + «3ffl , (9-4d)


STRESS 93

stress s in the oblique section plotted as a vector from a fixed point, s


being the radius vector
s = vV + s„2 + s,2 (9-6c)

in the stress ellipsoid Eq. (9-6b).


Using Eqs. (9-6) and (9-3) we obtain an expression for the shearing
stress t in the oblique section :

r2 = <rSax* + a fa* + <r35a22


- (^a*2 + otfj + <r3a?y . (9-7)
CHAPTER 10

MOHR'S REPRESENTATION OF STRESS


10-1. Plane Stress. Stress Circle.
We obtain a state of plane stress
in the xy plane if we take the normal stress a, = 0 as well as both shearing
stresses r„ = 0, t,„ = 0. Consider a prismatic element (Fig. 10-1)
formed by two planes parallel to the xy plane, one parallel to the zx, and

IT

ISI c
— V .
*
a,:
I
V

Fio. 10-1. Plane stress. Fio. 10-2. Mohr's stress circle.

one parallel to the zy plane together with an oblique plane parallel to the
zaxis. Denote by a the angle between the normal n to the oblique plane
and the x axis, by <rx,ay,r„ = ry, the stress components in th« two per
pendicular planes, and by <r and t the normal and shearing stress compo
nent in the oblique plane. The equilibrium of the forces acting on the
prism is expressed by the equations:
a cos a — t sin a = <r, cos a + tv, sin a ,
(10-1)
<r sin a + t cos a = <rv sin a + tw cos a .

Solving for <r and r we obtain

<r = —T +
i ———°V
<Tx
' Cos
n i
Tiw Sm

'
2
(10-2)
sin 2a + cos 2a .

By transposing the term (<r, + <r„)/2 in the first equation to the left side,
squaring, and adding there results

— +r +T«" (10-3)
\~2~
V1"

in the variables a and an equation of the "stress circle" of state


of
a
t

plane stress. By plotting a as abscissa and as ordinate (Fig. 10-2) in


t

94
MOHR'S REPRESENTATION OF STRESS 95

rectangular coordinates a, r the normal stress a is seen to reach its extreme


values a\ and a2 at the points Si and St of the stress circle. For these
I
da da = 0 and

tan 2a' =
2t"
(10-4)

while the shearing stress t reaches its extreme values at the points where
drfda = 0 and
tan 2a" = - ffx
~
"v
(10-5)

showing that
tan 2a' tan 2a" = - 1 , (10-6)
that the angles 2a' and 2a" must differ by a right angle and the shearing
stresses reach their extreme values in sections which make angles of
45 deg with the axes of principal stress. From Eq. (10-3), by taking
t = 0 we obtain the extreme values of the normal stress er or the principal
stresses

Also the absolute value of the largest shearing stress is

T°" =
\\~2 / + Tl* w._fi-«l. 2
(10-8)

We may express the Eqs. (10-2) and (10-3) by choosing the principal
directions as the coordinate axes. If a denotes the angle between the
normal n and the principal stress <n we obtain

ff=ZL+^ + kL^cos2a, T=-^!^sin2a. (10-9)

Conversely, if the angle a which the algebraically larger principal stress


<n( > oi) makes with the x axis, and both principal stresses are given, the
components <r„ <r„ and t„ may be computed from

~
«•*, *« o — o cos ^a ' Tl» = o Sln *a (10-9a)

and for Mohr's circle:

For example, for a state of pure shear <n


= -\-<r0, oi — —
<r0 we have

a = +<7o cos 2a ,
,, (10-10)
t = —
<TB sin 2a , 'i
90 THEORY OF FLOW AND FRACTURE OF SOLIDS

10-2. Mohr's Graphical Representation of a State of Stress.1 Con


sider a tetrahedron of material under a homogeneous state of stress, in
which three mutually perpendicular faces are so oriented that they
coincide with the principal planes of stress which are also the coordinate
planes yz, zx, and xy. Denote by 01,02,03 the principal stresses and by
at,ay,at the direction cosines of the normal to the fourth, oblique face of
the tetrahedron. This oblique
plane {ABC in Fig. 10-4) in space
may also represent a tangent plane
of a unit sphere, namely, the plane
in the first quadrant which is nor
mal to the radius OP of the sphere
in a point P, this point having the
rectangular coordinates equal to
the direction cosines ax,Oy,a, (Fig.
10-5). As this direction changes,
point P of tangency moves on the
surface of the unit sphere:

af + a,1 + a,2 = 1 (10-11)

carrying with it the plane of tan


gency. While point P moves on
the surface of the unit sphere con
sider a second point Q related to P
moving in a plane. Point Q has
the rectangular coordinates: ab
scissa a, ordinate t. a and t denote
F10. 10-3. Otto Mohr (from a book dedi
cated by his former pupils on his eightieth
the normal and the shearing stress
birthday, 1916), W. Ernst und Sohn, Berlin, in the oblique face of the tetrahe
(Courtesy Prof. Otto Graf, Stuttgart.)
dron [or in the plane of tangency
of the unit sphere, Eq. (10-11)]. az,av,a, are also the direction cosines
of its normal.
The condition of equilibrium of the forces acting on the tetrahedron
furnishes the values for the normal stress a and the shearing stress t in
the oblique face. Suppose that a normal stress a% alone would act across
the face of the tetrahedron coinciding with the xy plane. This would
1 Mohr, Otto, " Abhandlungen aus dem Gebiete der technisehen Mechanik,"
2d ed., pp. 192-235, Berlin, W. Ernst und Sohn, 1914. In these collected works by
Mohr he has published in revised form his most important papers. The one on
pp. 192-235 reproduces Mohr's original paper, published in Zivilingenieur, 1882,
p. 113, in which he first described his graphical representation of the general state of
stress. A second revised form of this paper was published in Z. Ver. deut. Ing., 1900,
p. 1524.
MOHR'S REPRESENTATION OF STRESS 97

produce in the oblique face a resultant stress s, in the direction parallel


to the positive z axis which is equal to sz = o-3a,; hence in the same plane
a normal stress a = 03a,1 and a shearing stress t2 = s,2 — <r2 = ffa'a.2 — <r2
must act. If all three normal stresses a\,<a,a% act simultaneously, the
normal stress must be equal to

a = <riOx2 + ff2a»2 + tfsO.2 • (10-12)

For the shearing stress r in the ob


lique plane we must obtain
* = **
s - ,T» = r2 + S„2 + S.1
-
= ff^Ot2 +
+ <r32a,2
-
<722a»2

a2 (10-13)

which are the Eqs. (9-5) and (9-7).


These two equations, (10-12) and
Fig. 10-4. Tetrahedron.
(10-13), may be interpreted also by
stating that through them a relation is established between two points P
and Q. The first point P(ax,av,ax) is situated on the surface of a unit
sphere. The second Q(a,r) is assumed in a plane, which we may call
"Mohr's stress plane o,t." We may describe the properties of a general
state of stress by studying the relation and the shape of certain curves

Fio. 10-5. The tangent plane to unit Fig. 10-6. Graphical representation of
sphere in P defines orientation of oblique normal and shearing stresses a and r.
section.

which may be traced, for example, on the unit sphere and of their images
in the stress plane prescribed by the Eqs. (10-12) and (10-13). This is
facilitated by considering three pairs of poles such as correspond to the
points Pi,Pi,Pi on the sphere (Fig. 10-5), the equators, parallel circles,
and meridians corresponding to each -pair of these poles and their images
in the a, r plane of point Q (Fig. 10-6).
98 THEORY OF FLOW AND FRACTURE OF SOLIDS

Solving the three equations (10-11), (10-12), and (10-13) for oI,,a»,fo,*
we obtain
— — — —
2 = (^2 <r)(<r3 a) + t2 „ — (<r3 <r)(<ri a) +T2
a — — ' " — —
(<r2 <ri)(<r3 ai) (<r3 <r2)(<ri a2)
— — t* (10-14)

<r)

+
(ai <r)(<T2
<V = — —
(<ri <T3)(<r2 <r3)

The images Qi, of the three poles Pi,P2,P3 of the unit sphere
Q2, and Q3
are found on the a axis of the stress plane, since for them = and

0
t
a = <ri, = <r2, = <rj, respectively, from Eqs. (10-12).
we hold, for example, a, = const, point moves along a "parallel

A P
If

circle" around the pole P3 on the sphere. curve, whose equation

is
obtained from (10-14), corresponds in the "stress plane" <t,t to this circle
we hold a, = const in the third of the three expressions. We obtain
if

— —
+ t2 = a,2(<r\ — — = const
a)

(<ri <r)(<r2 <r3){<ri <r2)

,
which the equation of a circle in the variables Or
is

<r,t.

Iff + = asiai - — = const .


t

<r3)(<r2 <r3)

j
)
2

The image curves of the "parallel" circles are family of concentric


circles. The image of the equator a, = has the equation a
0

- '—^y t2 =
^r^' -
a(M5)
+
(*

The image circle of this equator of the unit sphere and the image circles
of all parallel circles corresponding to in the <r,t plane have the same
it

center situated on the a axis in the distance (<ri 4- <r2)/2 from the origin

0
(point M\i in Fig. 10-9). Equation (10-15) defines one of the three
"principal Mohr stress circles" in the <t,t plane. Its radius equal to
is

(<ri

<r2)/2. Three principal Mohr's circles correspond to the three
"equators" a, = = = of the sphere. They pass through
0,
0,

Oy a„
two of the three poles Qi,Q3,Qi, respectively. By dividing a,2 by a„J
taken from Eqs. (10-14) one obtains similarly in the <r,r plane the equation
of a circle when this ratio a,2/Oy2 = const. But a,/Oy = const defines
a

meridian circle on the sphere. Therefore, the images of all meridian


circles traced on the sphere are also circles in the <r,t plane. The centers
of the images of the meridian circles in the <t,t plane are located on the
a axis.1

Figures 10-7 and 10-8 show the first octant of the unit sphere in two projections
1

with the families of parallel circles and meridians belonging to the pole P%. The
MOHR'S REPRESENTATION OF STRESS 99

Figs. 1,-7 and 1,-8. Flo. 10-9.

10-3. Shearing Stress r. Principal Shearing Stresses. Octahedral


Stresses. For the shearing stress r from Eq. (10-13) we obtain

t2 = <ri*a,2 + + "s^a,2 ~ (<ria,2 + <rW +


a22a„2 <r^a,2)2 (10-16)

or also

t2 = (a, - <r2)*a,V + (<r2


- <r2)W<h* -(-
- <r,) Va*2 . (10-17)

This equation defines the magnitude of the shearing stress t in the


oblique section. Along each principal stress circle we obtain a maximum

angle 7 defines the latitude and the two complementary angles i' and 5" the longi
tudes, which are marked in degrees along the meridians. Since the three principal
circles in the stress plane are Mohr's circles of the plane stress distributions at, <n; a?, a3;
a2, <ri, the simple angular rules for locating the corresponding points in which the
meridian circles intersect the equator apply also to Fig. 10-9. The angles y, S', and S"
are expressed in degrees. The points M are the center points of the principal stress
circles, M 12 is the center of the image circles representing the parallel circles y = const
of the unit sphere, etc. The image circles representing the meridians on the unit
sphere in the a, t Mohr plane are easily constructed. They pass through a pole and
through the point in which they intersect the corresponding equator and their center
points are therefore also easily located on the <r axis.
100 THEORY OF FLOW AND FRACTURE OF SOLIDS

= 0
(and minimum) value of the shearing stress t; for example, when a,
we have t2 = (a — <r2)2ax2a„2. Since in the x,y plane ax = cos a,
Oy = sin a we see that the stress

r = ± ^-^ sin la (10-18)

becomes an extreme value r = — when a = 45 or 135 deg.

cr2|
|ci

2
:
Summing up we see that the middle points of the three principal stress
circles have the distances from the origin of the stress plane q,t.

0
IL+*?, 0+^,21+*! (1(M9)

and that the absolute values of the principal stress differences divided by
two represent the radii of these circles:

Ol —
— —
<7"l ff2 <7"3 <T"3 G\
(10-20)
2~~

'
'
2

2
Let us define by

tj ——
= ffi <r2

<r2
— <r% —
_<ri
— Q\
/,nQn
Tl = (10-21)

2^
'
g

'

in the cyclic order 1,2,3. These three stress components ti,t2,t3 are
called the principal shearing stresses of the state of stress. They satisfy
the equation
Tl + =
+

. (10-22)
0

T2 T3

This convention defines the rule according to which the positive direc
tions for the principal shearing stresses ti,t2,t3 are to be chosen. They
are to be taken positive when, in the plane stress distributions 03,01; tri,<r2;
<r2,<r3, the arrow of the shearing stress points in the direction from Pi to P3,

Pi to Pi, and P3 to P2 in cyclic order. According to this convention the


sign of one of the three principal shearing stresses must always be negative,
the two others were positive.
if

we place cube in such a way that its sides coincide with the planes
If

of principal stress, the sections in which ti,t2,t3 act form rhombic


a

dodecahedron. The directions of the principal shearing stresses form the


edges of a regular octahedron whose corners lie upon the principal axes.
!

Using ti, t2, and t3 the shearing stress in any oblique plane may be
r

expressed by
t2 = 4(T32aI2ov2 + nWa,2
+

T22a,2a,2) (10-23)
.

The positive directions for the principal shearing stresses are indicated by the
1

arrows of the vectors, ti, T2, t% in Fig. 10-10.


MOHR'S REPRESENTATION OF STRESS 101

A rule may also be helpful for determining the direction in which t acts in this
plane. For this purpose let us resolve the vector r in its plane into the three compo
nents in the direction of the three merid
ian circles of the unit sphere which inter
sect in the point P(aIta„a.). Suppose
that only one principal stress, e.g., a%, acts.
The traction <r» produces in the triangle
ABC (Fig. 10-11) a resultant stress

S, = ffj cos 7

and this has a component t3„ in the plane


ABC which is

ti» = s, sin 7 = (<r, sin 27): 2.

This component tj„ acts in the meridian


plane passing through P and the pole P>; in Fig. 10-10. The planes of the principal
the triangle ABC it acts along the height shearing stresses define a rhombic dodec
CC Thus we see at once that the shear ahedron.
ing stress may be resolved into three com
ponent vectors in the plane of the triangle ABC if a\, <n, and o3 act; each must be
perpendicular to one of the three sides of the triangle ABC. This has been shown

NORMAL n

Fio. 10-11. Oblique section.

in a plane view of the triangle (Fig. 10-12), in which these three components were
traced from the center of gravity G of the rectangle. They are
sin 2a sin 20 sin 27
(10-24)

The components ti„ij„,ti, are taken positive in the direction in which they would art
if the principal stresses <n,o-2,o-a were positive. The resultant of these three coplanar
102 THEORY OF FLOW AND FRACTURE OF SOLIDS

vectors must be equal to r. By trigonometry the cosines of the angles Si, Sj, and Si
shown in Fig. 10-12 are found to be equal to

cos 81 = — cotan 0 cotan y ,


cos Sj = — cotan y cotan a , (10-25)
cos 63 = — cotan a cotan /3 ,

The direction of the resultant shearing stress t is thus known from its components and
their directions.1

cos A C cos a
siNf ""r"
cos/3 smr"
-/nr.
yaj dj
cosr
cos a cos/3
Fio. 10-12. Fio. 1,-13. Fio. 1,-14.

If we choose in the tetrahedron the direction cosines of the normal


to the oblique plane o, = ay = a, = l/\/3
= 0.577 by making the angles
a = /3 = y = 54°45', we obtain for the normal stress the value from
Eq. (10-12):
<f\ + <r2 + <TS
<tool — (10-26)

which is equal to the mean value of the principal stresses. For the shear-
1 The three plane views of the triangle ABC
which are shown in the Figs. 10-12 to
10-14 indicate the components of the shearing stress r in the direction of the meridian
planes (heights of triangle) and also in the direction of the parallel circles (the compo
nents now appear parallel to the sides of the triangle). The central figure may be
helpful when the angles 5i,i2, 53, heights, etc., are to be computed in ABC. Point P
in Fig. 10-11 is the point in which the normal drawn from the origin O to the oblique
plane of the tetrahedron intersects this plane. The distance OP = 1.
MOHR'S REPRESENTATION OF STRESS 103

ing stress t we obtain in the same plane from Eq. (10-17) the value

Toot = H V(<n
- «%Y + (»«
- <r»)2 + (ffi - <n)2 (10-27)

or from Eq. (10-23) also

t«, = % Vu* + r2* + t,2. (10-27o)

There are four pairs of parallel planes, which have for their normals the
directions ax = au = a, = ±l/\/3-
As tangent planes of the unit sphere
they form a regular octahedron in
space whose corners lie on the prin
cipal axes (Fig. 10-15). ffxt and T„ct
are called the octahedral normal and
shearing stress, respectively. We
shall seelater that the octahedral <£

shearing stress TMt measures the in


tensity of stress which is responsible
for bringing a solid substance into
the plastic state.1
If the principal directions are not
known and the state of stress is given
Fio. 10-15. Octahedral planes of stress.
by the six components of stress ax,

a*, at, Txy, Tyz, t«, and Toct


<7<>ct are expressed as follows:

<Tx + <rv +
(10-276)

t.,=H V(^-^)2+(^-^)2+(^-02+<Kt^2+tvj2 2). (10-27c)2


1 The meridians passing through point Pt for an octahedral plane bisect the angle
between the principal planes of stress (Figs. 10-16 and 10-17). The image point Q0
of P0 in the plane a, r having the coordinates o-oot and Toot is shown in Fig. 10-17.
[ffoei being equal to the mean normal stress (o-i + o-2 + a j)/3 is denoted by a and Toct
by to in Figs. 10-17 and 10-18.]
* Since
o-oct and Toot depend on the three principal stresses <n, at, o-3 their expressions
given by Eqs. (10-26) and (10-27) must be invariants of the state of stress. The sum
»» + «■» + c« appearing in Eq. (10-27b) is indeed equal to the "first" invariant
•"i + at + cj of the tensor of stress (see Chap. 9, p. 92). The second expression
given by Eq. (10-27c) for the octahedral shearing stress Toot is obtained by noting that
the right side of Eq. (10-27) depends on both the "first" and "second" invariants of
the state of stress and may be expressed also by

Toot = % V(o-l + <T! + O'j)*



3(0"10*| + O-jO-j + a3B\) .

After substituting the expressions, which were given in Eq. (9-4c), for the values of
the first and second invariants appearing in the two brackets in the last equation
104 THEORY OF FLOW AND FRACTURE OF SOLIDS

The magnitude of and the direction in which the octahedral shearing


stress T„ot acts can be constructed very simply without knowledge of the

Flo. 10-16. Fig. 10-17.


Flos. 10-16 and 1,-17. Octahedral plane and stresses ero and to.

preceding if the principal stresses are known. Noting that the triangle
ABC in an octahedral plane is an equilateral one, if we choose OA = 1
(Fig. 10-11), the triangle has the
sides y/2 and the area \/3/2. The
stress component s, balancing the
principal stress <r3 is s, = <r3/y/Z.
It contributes the component
Un = V% 8, = y/2 <r3/3
to the stress Toot along a height of the
triangle ABC. The stress To*t must
therefore be the resultant of the
three coplanar components y/2 ,ri/3,
y/2 a2/3, and \/2 <r3/3 each of which
acts along a height of the equilat
Fio. 1,-18. Octahedral shearing stress to.
eral triangle ABC and which must
make equal angles of 120 deg with each other (Fig. 10-18). If a vector v
has three components Vi,v2,v3 in one plane, which are inclined with respect
to each other at 120 deg, one verifies easily that

f2 = MKfi - »2)2 + to - v3Y + (», - Vly] (io-27d)

Eq. (10-27c) above is obtained. (See v. Mises, R., Goeilinger Nachr., 1913). The
fact that in the plasticity condition of the ductile metals [see Eq. (15-18)] the second
invariant has a prominent role has been unduly emphasized by some investigators.
Any expression containing the three principal stresses is an invariant of the state of
stress, i.e., independent of the choice of the coordinate system.
MOHR'S REPRESENTATION OF STRESS 105

Inserting here for Vi = V 2 ci/3, . . . gives for v the expression derived


above for the octahedral shearing stress T«t in Eq. (10-27). These three
components define finally the direction of their resultant, namely, T„,t.

Fig.10-19. Fig. 10-20. Fig. 10-21.


Fio. 10-19. Planes of principal shearing stress. Fig. 10-20. Octahedral planes. Fig.
10-21. Planes of principal stress.

10-4. Deviator of Stress. A homogeneous state of stress is called a


deviator of stress if the sum of the principal stresses vanishes. The gen
eral state of stress can be split into the sum of a uniform tension or com
pression, equal in all directions, and a deviator. If 0-1,0-2,0-3 are the
principal stresses and their mean value a = (<7i + o-2 + o-3)/3 vanishes

or if
<Tl + <T2 + 0-3
= 0 (10-28)
they are the principal stresses of a deviator of stress.
Deviators are of importance in the theory of plasticity because the states
of stress causing flow in one of the most important classes of formable
materials, namely, in the ductile metals are essentially deviators. The
distributions of stress under which the ductile metals start to yield and
under which they continue to deform permanently do not seem to depend
appreciably on the mean stress a. Deviators are also encountered in the
states of strain of incompressible materials, as we shall see later.
106 THEORY OF FLOW AND FRACTURE OF SOLIDS

This experimental fact may be expressed by stating that the figure


consisting of the three principal Mohr's stress circles determining a state
of stress at a limit of plasticity may be displaced or shifted as a whole in
the direction of the a axis in Mohr's stress plane without ceasing to repre
sent such a limiting state of stress. In other words, as long as the con
dition of plasticity does not depend on the mean stress (<n + <n + <rs)/3,
we may describe the entire range of all possible states of stress causing
flow by confining ourselves to the consideration of deviators of stress only.
Having such applications later in mind we
have to choose the origin O in a <r,r plane, in
|\^*— T">w y
1
\
which we wish to represent a deviator of
stress through its principal circles at the cen
ter of gravity of three equal masses, which
would be placed at the points Qi, Q2, and
Q% along the a axis. (A construction for
locating the origin O of a deviator is indicated
in Fig. 10-22.) It is customary to assume
that algebraically <ri > a2 > a3 (ai largest,
<rj intermediate, a3 smallest principal stress).

Fio. 10-22. Deviator. To define the position of point Q2 relative to


Qi and Q3 in Mohr's stress plane it suffices
to assume that the ratio n of the distance MQi to the radius
MQx = T = (<rx- a3)/2
of the largest principal circle is known:

= MQ2 = 2<ri
- - ax <r,
M XfTT — a3 (10-29)
MQi ai
where M denotes the position of the center of this circle. The principal
stresses of any deviator can then be expressed as follows:

^ =
(l-|)r, =
»3=-(l+^r (10-30)

The distance MO = nr/3. A deviator of stress is thus determined by


two quantities only (jt, t) if the direction in which the principal stresses
act are known and all deviators of stress are obtained by varying n
between
-l^/igl. (10-31)

Three special deviators shall be distinguished according to:

When Qa coincides with Qs: <r%


= a3 , n = —I .

When Qi coincides with M: a2 = (<ri + a3)/2 = 0 , n = 0 .

When Qi coincides with Qx: a2 = <r\ , n = 1 .


MOHR'S REPRESENTATION OF STRESS 107

The case n = — 1 characterizes a generalized


1 a tensile and n =
generalized compression test (since the corresponding Mohr's circles are
the same as for the ordinary tensile or compression test after the origin
is moved from its location in the deviator to points Q3 or Qi). p = 0
describes the state of a pure shear.

/i'0 ji'l
Fio. 10-23. Fio. 10-24. Fio. 10-25.
Fig. 10-23. m ■ — 1. Generalised tension test. Flo. 10-24. n — 0. Pure shear test.
Fig. 10-25. ii ■ 1. Generalised compression test.

The constituent deviator part of a general state of stress (01,02,03) may


symbolically be abbreviated by the schemes:

0\ — 0 0 0
V. = 0 at — a 0
0 0 — a
03

H
— a Ty* r,x
0x
— 0
Tzv 0y Tzy (10-32)
TIZ Ty:
— O
oz

H. Hencky emphasized the usefulness of this nomenclature in dealing


with the plastic states in materials. A. A. Ilyushin1 went one step
further by proposing to utilize a "deviator of similitude" (of stress).
Its components are obtained by dividing the components of stress by the
value of the octahedral shearing stress

to = H Wi —
o^)2 + •

This "unit" deviator is given by

0\ — 0
0 0
To
— a
0%
D. 0 0
To

0-3
-- 0
0 0
To

1 See footnote, p. 85.


108 THEORY OF FLOW AND FRACTURE OF SOLIDS

The octahedral shearing stress of the deviator of similitude is equal to


unity and remains unity for all plastic states.
The experimental work required for determining the rules under which
the ductile metals start to deform plastically may be guided by the pre
ceding remarks. Tests should primarily be concerned with measure
ments facilitating the evaluation of the radius r of Mohr's largest prin
cipal stress circle for a few characteristic values of n such as, for example,
n = —1,0,1. If then the mean stress a were also varied, this would per
mit a survey of the entire field of stress.
A state of pure shear is a deviator (a\ = — a2,<Ti = 0). A deviator of
stress can be resolved into the sum of two constituent states of pure
shear:

Since the deviatorial part <r\



a, a a, ttj — a

of a general state of stress having the principal stresses <ri,<r2,<rg may be


resolved as follows:
— a
a\

— a
a2

— a
a3

or
— a =
<ri

— —
(<rl <r2) <ri a2
1 " '
3 3

a3
— a = 0 - (<ri

3
<r2) {a2

3
<r3)

it is seen that a deviator of stress may also be split into three states of
pure shear appearing in the three columns of terms on the right sides of
the last three equations.
CHAPTER 11

STRAIN

11-1. Homogeneous Strain. If we wish to determine the change in


shape of a body, we have to compare the positions of its material points
in the original state with their positions in the second state. Thus a
point P (Fig. 11-1) is displaced from its initial position with the rectangu
lar coordinates x,y,z to some point P' having the coordinates x',y',z'.
These latter as well as the components of the
line PP' parallel to the coordinate axes or the
components of displacement

u = x' — x , v = y' -y , *4.


of the point P are in general functions of the
coordinates x,y,z and the time. A*
The simplest cases of deformed bodies are , /
those in which the displacements u,v,w are linear
*\
!/ /
functions of the coordinates x,y,z. We may XJ
leave out the constant terms in these functions „ , , ,
Fio. 11-1.
since they represent only a parallel displace
ment of the body as a rigid body and we may assume that these func
tions are homogeneous. In this kind of deformation, all points, which
initially lie in a plane, after distortion lie again upon a plane; more
over, parallel planes after distortion remain parallel. The point coincid
ing with the origin O of the coordinate system x,y,z remains at rest.
These kinds of distortion are known as states of homogeneous deformation
or strain. Simple examples are the following:
a. Pure Extension. If all points are displaced parallel to the x axis,
the displacements u being proportional to x and v = w = 0, we have
u = cx , v = 0 , w = 0 .

The constant c If 0 < c the body is elongated,


must be larger than — 1.

if — 1
< c < 0 compressed in the direction parallel to the x axis. The
measure of the extension of the body is the extension of the unit length
or the unit elongation c.
b. Simple Shear. If the points are displaced in a direction parallel to
the i axis through distances proportional to y:
u = cy , v = 0 , w = 0
109
110 THEORY OF FLOW AND FRACTURE OF SOLIDS

the deformation is called a simple shear. In this case all planes parallel
to the xz plane slide in the direction x without changing their distances
from each other and the displacements are proportional to their distance
from the xz plane. The measure of shear is the unit shear c, or the dis
tance by which two of these planes, of unit distance apart, are displaced
with respect to each other.
Two families of straight lines
under this state of strains do not
change their lengths, namely, the
straight lines y = const and the
Fig. 11-2. Fio. 11-3. Fio. 11-4. straight lines y = — (2x/c) + const,
Fios. 11-2 11-3, and 11-4. Simple shear.
Left — initial, right — final state of strain. as shown in Figs. 11-3 and 11-4.
This type of deformation is well
known in the case of the plastic distortion of metal single crystals. The
formation of a twin in calcite is another example for it.
c. Uniform Dilatation.

u = ex , v = cy , w = cz .

d. Longitudinal Extension without Change in Volume,

u = ex , v = —c'y , w = —c'z , c' = 1 -


vT+c
e. Pure Shear.

u ex v = — y w = 0
1 +c

§
This deformation consists of a uniform extension in the x direction and a
uniform contraction in the y direc
tion of such an amount that the B /ik
volume remains unchanged. In
,/ s
/ ^
A'y '
/
this case a certain rhombus A BCD
(Fig. 1 1-5) is distorted under strain v. s_ X
into a congruent rhombus A 'B'C'D' N
(Fig. 11-6), in which the acute and \D'
D
the obtuse angles have been inter Fio. 11-5. Fia. 11-6.
changed. Fios. 11-5 and 11-6. Pure shear. Left-
initial, right — final state of strain.
/. Pure Rotation. Homogeneous
linear functions of the coordinates x,y,z may also express rotations of rigid
bodies without strain. For example, the function
x' = x cos a — y sin u ,

y' = x sin <o + y cos to ,

where to is a given angle expresses a rotation of a rigid body around the


STRAIN 1 11

z axis by an angle «, since r*' = x'2 + y'2 becomes equal to r2 = x2 + y2.


The displacements m,i>,u> are expressed by the equations:

u = — (1 — cos o>)x — y sin u> ,

v = x sin w — (1 — cos w)y ,

w = 0,

from which we may note that a pure rotation (without any distortion)
around the z axis by a finite angle u may be constructed by the simul
taneous superposition of three strains, namely:

A simple shear: u = — y sin w , v = w = 0 ,

A simple shear: u = w = 0 , «; = x sin o> ,

A uniform dilatation: u = —
(1
— cos u)x , v = — (1 — cos w)y ,

w = 0.

Although in some of the preceding examples simple strains were suc


cessively combined with each other by adding the corresponding terms
in the equations it is in order to note that the commutative principle of
addition does not hold for finite homogeneous strains. This is well
known of finite rotations of rigid bodies; the order in which two finite
rotations of a body are added to each other is essential and cannot be
reversed. The same rule must be strictly observed if two or more finite
strains are carried out in succession in general. Although in certain
examples the commutative principle may be valid, in others it is certainly
not applicable.1
Suppose that the coordinates x',y',z' of a point in the strained state of
the body are linear homogeneous functions of the coordinates x,y,z of
the same material point in the unstrained condition:

X1 = CUX + C12J/+ C13Z , \


y' = C21X + c22y + c23z , (11-1)
I
z' = c3ix + dty + C33Z . )

A set of equations like (11-1) represents what is known as a linear trans


formation in space.
Consider a sphere inscribed in the body, its center being at the origin 0
and its radius being equal to r while the body is in its initial unstrained

1 For example, two simple shears, the planes of which coincide, may be added as
vectors geometrically and their components by simple addition and in any order.
The order, however, in which two finite simple shears are added, which do not act in
in the same planes, does not follow the commutative rule.
112 THEORY OF FLOW AND FRACTURE OF SOLIDS

condition. The points on this sphere satisfy the equation:

+ + z2 =

j/2

r2
x2 (11-2)

The deformation or the linear transformation given by the Eqs. (11-1)


may be visualized by the surface into which the sphere given by Eq. 1-2)

(1
has been distorted through the strain. The equation of this surface is
obtained by substituting in Eq. (11-2) the values of x,y,z obtained by
solving the three linear equations (11-1) for x,y,z. function of the

A
second degree in x',y',z' results, provided that the determinant of (11-1)
does not vanish. Since x',y',z' according to (11-1) are finite, x,y,z have

if
finite values and x,y,z must be finite be
cause they satisfy Eq. (11-2), the surface
must be an ellipsoid. Its middle point

is
the point 0. Consider a cube circum
scribed about the sphere in the unstrained
body. Under strain this cube becomes
an oblique-angled parallelepiped tangent
to the ellipsoid while the three mutually
perpendicular diameters of the sphere
which connected the points of tangency
result in three conjugate diameters of the
jA

JKt ellipsoid. Since this must also hold for


^ ^ose three mutually perpendicular diam-
*9 jf^^L JmBS
*
T

4Hp eters of the sphere, which after distortion


<j9

become the principal axes of the ellipsoid,


i.e., which after distortion again are mutu
ally perpendicular, may be recognized
it

Fig. 1-7. Augustin Louis Cauchy


1

at once that homogeneous strain defined


a

(1789-1857).
by the Eqs. (11-1) may be divided into:
(1) three pure extensions of the body along three mutually perpendicular
directions and combined with (2) rotation of the body such as to bring
a

these three directions into coincidence with three other mutually perpen
dicular straight lines, namely, the principal axes of the ellipsoid. The unit
extensions along the principal axes of the ellipsoid are called the prin
cipal strains. Since a rotation of rigid body about fixed point
is
a
a

described by three parameters while Eqs. (11-1) contain nine constants,


we see that the case of pure strain without rotation described by six
is
a

= the ellipsoid called the strain ellipsoid. If this


If

quantities.
is
r

becomes sphere congruent to the sphere given by Eq. (11-2), the body
a

has only been rotated without any strain.


The most general expression describing state of pure homogeneous
a

strain without rotation defined by three linear equations


is
STRAIN 113

y + 2 .

(11-3)
c,, I ^if, ,

in which the coefficients dy = cyx, cv, = c,y and c,, = c,, (c/. also Chaps. 12
and 14). Equations (11-3) define what is known as a symmetric linear
transformation.1 The three pure extensions u = c,,x, v = cmy, w = c,,z
indeed do not produce a rotation of the principal
axes of strain and the same is true of any of the
three shearing displacements of the form (see
Fig. 11-8):

" =
cy,
"2
V> V
-f x w = 0, (11-4)

provided that c„ = cyx, etc. Under the state of


strain defined by Eqs. (11-4) the straight lines
making an angle of 45 and 135 deg with the x axis remain parallel to these
two mutually perpendicular directions, and hence the principal directions
of strain coinciding with these directions do not rotate for any of the three
similar types of shearing distortions.
11-2. Infinitesimal Strain. In accordance with engineering practice
we shall henceforth denote a unit extension, i.e., the ratio of the increase
of length to the original length by the symbol « and affix to the unit
extensions [the normal strains and coefficients which appeared in the
terms on one of the diagonals in the linear transformations (11-1) or
(11-3)] only one subscript. This subscript will usually indicate the direc
tion a line had in the unstrained state of the body. Unit shearing strains,
in accordance to their definition in Sec. 11-1, will be denoted by y attach
ing, in general, two subscripts to the letter y, e.g., y^. When the shears
are small the first subscript may denote the direction of the normal
to a plane which participates in a sliding motion relative to planes
parallel to it and the second subscript may indicate the direction of
the slip.
Rewriting Eqs. (11-3) for a pure strain not containing a rotation in the
symmetric form:

1 The sixconstants appearing in Eqs. (11-3) are subject to restrictions. The values
of c,x, c„y, c„ must be algebraically larger than —1 and the remaining coefficients
cxy, cy,, c„ must be — 2 < c,y < 2, etc.
114 THEORY OF FLOW AND FRACTURE OF SOLIDS

u y +
4- — z '
2 2

v
(11-5)

w x + y + «,z >
~2 "J"

in order that they shall represent a state of infinitesimal pure strain the
three components of displacement u,v,w must remain small quantities
in comparison with the coordinates x,y,z. The constants t,t . . . 7„,
. . . must therefore be small relative to unity. The constants e„,«„,«,
are the unit extensions in the direction of the axes x,y,z and = 7^
7„, = fn, 7« = 7„, are the
unit shears.
We may note that the three rectangular components u,v,w of a displace
ment vector v in Eqs. (11-5) are represented by three linear expressions
of the three rectangular components x,y,z of a space vector r in a way
similar to that in which the three components s,,sy,s, of the resultant stress
vector s in Eqs. (9-1) were expressed as dependent on the three direction
cosines a,,ay,a, (i.e., on the components of a unit vector a in space). We
may add that, since these two groups of equations (representing two
"linear vector functions") must have identical properties, a close sim
ilarity must exist between the properties of a state of infinitesimal strain
expressed by Eqs. (11-5) and those valid for a state of homogeneous^stress
defined by Eqs. (9-1). To make the correspondence perfect, suppose
that we circumscribe a sphere around the origin O so that

x2 + y2 + zi = r* .

If we divide Eqs. (11-5) by r and choose r = 1 = const we may substitute


in the Eqs. (11-5) for x/r = a,, y/r = a„, z/r = a, the direction cosines
a,,ay,a, of a unit vector a. u,v,w will then represent the small displace
ment components of the points x,y,z in a body which in its unstrained
condition were situated on a unit sphere. Because of the identical forms
of the right sides of the latter equations and of the three Eqs. (9-1), we
must therefore conclude that if we replace in the equations of a homogene
ous state of stress the six components of stress:

<T,, 0"]/, <T,j Tygj Ti,, ^xy (11-6)

by the six components of strain:


Ty, fi, 7i„
*x>
2
'
2
'
2
(11-7)

we obtain the corresponding laws for a state of infinitesimal strain.


STRAIN 115

Therefore, for example, Eqs. (11-1) are referred to the principal axes

if,
of strain, they become, taking as the principal strains,

u = tix = ny w = tjz . (11-8)

v
,

,
In any given direction (a,,ay,a,) the strain equal to

is
= a,2d + ay2t2 + ax2«a . (11-9)

e
The resultant unit shear in those planes which are perpendicular to

7
the direction {a,,ay,a,) given by

= aSoy^ti —
+ is
ay2aS(t2 — t%Y 0/ r-
-j

I
«2)2

!
-

I I
+ a,*a,*(«, «,)* (H-10)


If the direction perpendicular to the octa
if is

I
l/\/3,

ft
hedral planes, i.e., a, = a, = a, =

I
ox

±
we obtain the strain and unit shear:
Fio. 1-0. Infinitesimal

1
unit shear. 11-9In Fig.
«i + «2 + tj the more general case of an
infinitesimal strain involv
ing an infinitesimal rota

+ («, - «,)'] (li-ii)i


tion by an angle
fb% du\

1
Us
\dx dy)
A

2
state of infinitesimal strain represented
is

around the axis is as

2
graphically by its three principal strain circles in sumed. For pure strain

a
a manner similar to Mohr's representation of without a rotation (the case
a

considered in Fig. 11-8)


state of stress (the ordinates of the strain circles
dv du
must be taken equal to one-half of the unit u, =
0,

dx dy
shears y, however).
If the small components u,v,w of the displacement are functions of the
coordinates x,y,z, we proceed as follows for determining the expressions
for the components of strain. Consider the change in length and change
in angles of small element dx dy dz. length dx parallel to the x axis
A
a

changes under strain by an amount du/dx dx we proceed in the x direc


is if

tion. The unit extension in the x direction t, = du/dx. Correspond


ing to this tv = dv/dy, «, = dw/dz. The three unit extensions or strains
in the direction of the x,y,z axes are therefore

f* ~
du _ dv
u ~
dw
(11-12)
t"~dl,' dz
'

die

Furthermore, during deformation the right angle between the edges dx


and dy (Fig. 11-9) of an infinitely small parallelepiped changes by a
If the coordinate axes are not referred to the principal axes the octahedral unit
1

shear expressed by
is

70

- %V(u - «,)' + + Mr,.* + (ll-lla)



To
-)
116 THEORY OF FLOW AND FRACTURE OF SOLIDS

small value (du/dy + dv/dx). For infinitesimal strains this change in


angle is the unit shear 7„ = yyx. The unit shearing strains in the
material are therefore

du dw
-B + 5-
_ dv dt) ,
7l" ~ + ' 7v' ~ + ' <"-13>
di dx di dy~

The volume of this right-angled parallelepiped dx dy dz in the distorted

Fia. 11-10. Strain ellipses on compression test piece of paraffin. On one side of the speci
men circular scratches were cut by turning the face in a lathe. The ellipses shown in the
photograph were formed by the compression from the circular scratches. By this method
the axes of principal strain in a state of plane stress and the magnitude of the principal
strains can be determined experimentally.

condition is (1 +0(1
+ ev)(l + t,) dx dy dz. Since the strains are
small, the increase in volume per unit of volume or the cubical dilatation
is
du , dv . dw
«i + «i/ + U = -r
dx
r tdy r -r—
dz
(11-14)
CHAPTER 12

FINITE HOMOGENEOUS STRAIN WITHOUT ROTATION


12-1. Elongation X, Shearing Strain 7. Among the cases of finite
homogeneous strain we can eliminate the cases in which the volume of the
body changes through the remark that in such cases in addition to a pure
distortion without change in volume a uniform cubical dilatation « is
involved. The corresponding linear strain e0 may be computed at once
from the given volume dilatation t from the formula «o = v^l + «.
After subtracting from the terms of the linear transformation expressing
the general state of strain three equal terms representing the uniform
linear strain «0 we obtain a linear transformation without a change in
volume. In the applications when metals flow or rocks deform plastically
or rubber is stretched, the small changes of volume, which may be
observed by using very sensitive means of measuring the strains, can
usually be neglected. The consideration of an incompressible material
in these applications is sufficient, with a few exceptions in certain cases to
which later reference will be made.
A state of pure strain without rotation is obtained by stretching a body
simultaneously in three initially mutually perpendicular directions.1
These we choose as the axes x,y,z. Suppose that in the unstrained condi
tion a cube is given in the body with one corner of it coinciding with the
origin 0 of the coordinate system and with the three edges intersecting in
0 coinciding with the x,y,z axes. The sides of this cube shall have a
length r. After the strain is applied, the cube distorts into a brick-shaped
body. A sphere of radius r inscribed in the cube around the origin 0
is distorted in an ellipsoid. An octant of the sphere and ellipsoid is shown
in Fig. 12-1. A point P(x,y,z) situated on the sphere assumes in the
strained condition of the body a new position P'. Suppose that x',y',z'
are the coordinates of P', that ax,ay,a2 and aj ,aj,az' are the direction
cosines of the radii OP = r and OP' = r' of the sphere and ellipsoid,
respectively, so that

x = axr , y = <V , z = a,r , (12-1)


= az'r'
x' , V' = ay'r' , z' = a.'r' . (12-2)
If we designate by e1,€S,e3 the unit strains along the x,y,z directions and
1
Finite strains with rotation are treated in Sec. 13-4 and Chap. 14.
117
118 THEORY OF FLOW AND FRACTURE OF SOLIDS

by t the strain along the direction of OP', the coordinates x',y',z' of point
P' in the strained condition of the body are expressed by means of the
coordinates x,y,z by

x' = (1 + ti)x , y' = (1+ t,)y , z' = (1 + u)t (12-3)

and the unit strain t along OP' is equal to

(12-4)

Instead of the conventional strain t we shall use a new more convenient

Fio. 12-1. Finite strain.

variable X which we define as follows:

£
= + =
^ (1 «)2 (12-5)

and may call the (quadratic) elongation in the direction OP' in the body.
Similarly, if we introduce instead of the principal strains ei,«2,ej the
principal (quadratic) elongations Xi = (1 + «i)2, . . . the coordinates of
point P' may also be expressed by

= X,axV ,
= XjayV2 ,
= X2a,Jr2 ,
FINITE HOMOGENEOUS STRAIN WITHOUT ROTATION 119

and the elongation X in the direction OP' by

X - X,ax2 + X,av2 + X3a,2 . (12-7)

X can only be a positive number, X ^ 0.

Since the volume in the strained condition of the body remains un


changed, we have the condition that xyz = x'y'z' or

X^jX, = 1 . (12-8)

AAA
From Eqs. (12-2), (12-6), and (12-7) we express the direction cosines
az',at',a,' of radius OP' of the ellipsoid through the direction cosines
ax,Oy,a, of radius OP = r of the sphere as follows:

aJt = W ( an =
X^2 ^ aJt _ W .
(12_Q)

If the points P(x,y,z) lie on a unit sphere embedded in the unstrained


body
x2 + y2 + z2 = 1 (12-9a)

the points P'(x',y',z') in the strained condition of the body, according


to Eqs. (12-6) will be on the ellipsoid

r'
T>i v'2 -n
IT + X" + 1 (12-%)
Ai A2 Aj

which is called the strain ellipsoid. The lengths of its radius vectors are
OP' = r' = \/X.
Conversely, if a unit sphere is assumed in the strained condition of the
body (OP' = r' = 1),
x'2 + y'2 + z'2 = 1 (12-9c)

the corresponding points P(x,y,z) in its unstrained condition will be


found on the ellipsoid
Xix2 + X22/2 + \tz2 = 1 (12-9rf)

which is called the reciprocal strain ellipsoid. It has the property that
its radius vectors OP = r = l/\/\
are equal to the reciprocal values of
\/X for directions drawn in the undeformed state of the body.1
1 Love, A. E. H., "A Treatise on the Mathematical Theory of Elasticity," 2d ed.,
pp. 37, 60, Cambridge University Press, London, 1906.
The principal axes in both ellipsoids [Eqs. (12-96) and (12-9d)] coincide for a
homogeneous strain without rotation. For a strain with rotation they do not coincide
with each other, but the principal axes of the reciprocal strain ellipsoid define those
mutually perpendicular directions in the unstrained body which after the strain is
applied become the principal directions of strain, coinciding with the principal axes
of the strain ellipsoid.
120 THEORY OF FLOW AND FRACTURE OF SOLIDS

Equations (12-6) defining strain transform by means of Eqs.


a state of

(12-9) a pencil of rays {a,,ay,a,) originating from point O (Fig. 12-1) in


the unstrained condition of the body into a second pencil of straight lines
(fls ,aty' ,a,') embedded in the strained body. In order to simplify as much
as possible our notation in the discussion of these changes in orientation
of straight lines in space and of the changes in their lengths, we shall make
use of further abbreviations for expressing directions in space by writing
for
aS = a, V = b , a,2 = c . (12-10)

These parameters a,b,c we shall call the "directions" in our pencil of


rays. [They are essentially positive quantities smaller than or equal to
unity and could be interpreted, if one so desired, as the coordinates of the
points of an equilateral triangle situated in the plane

a+6 + c- l= 0 (12-11)

in an (a,b,c) system of rectangular coordinates.]


Equations (12-7) and (12-9), after making use of the abbreviations intro
duced in Eqs. (12-10), take the form

Xja + X26 + X3c (12-12)

V =
^ A
, c/ = ^A .
(12-13)

The last equations define the directions a',b',c' in the strained body which
originally were the directions a,b,c in the unstrained body.
Next we need an expression for a finite unit shear. Suppose that a
plane E normal to the direction OP(a,b,c) in point P is given. E is the
tangent plane of the sphere xS + + = r2 in point in the unstrained
z2
j/2

body. The plane after distortion becomes a plane E'. E' the
E

is

tangent plane of the ellipsoid shown in Fig. 12-1 in point P' (x',y',z').
Line OP which originally was perpendicular to plane in its new position
E

OP' inclined at certain angle with respect to the plane E'. The unit
is

shear in point P' defined by the tangent of the angle under which
is
y

\p

the line OP' inclined with respect to the normal n" to plane E'. unit
A
is

shear herewith defined in accordance with the previous practice


is
y

described in the examples of Chap. 11, page 109, as the slip of a second
plane parallel to E' one unit of length apart in the strained body relative
to plane E'\
- tan (12-14)
y

^
.
FINITE HOMOGENEOUS STRAIN WITHOUT ROTATION 121

This is visualized in Figs. 12-2 and OP = r is normal to plane E


12-3.
(Fig. 12-2) in the unstrained condition of the body and OP' indicates the
new position of OP in the strained body, similarly E' the position of the

fa, a. at)

(a; a-
o'y

\
/E' 33a7"~7i

Fig. 12-2. Fio. 12-3.


Figs. 12-2 and 12-3. Planes E and A".

points which originally were the material points of plane E. OQ' indi-
cates the normal to plane E' drawn from the origin 0.1
For the evaluation of the unit shear suppose that plane

E
is
y

represented by the triangle ABC


(Fig. 12-4). Xi = OA, Vi = OB,
If

z, = OC are the intercepts of plane


E and {,ij,f designate the running
coordinates of this plane, its equation
may be written in the form

- + - + --1=0
x, yi Zi
(12-15)

or in normal form
+ a,f =
+

az$ Oyij (12-16)


0
r

so that

- -A
r

az = a'
Xi
'
*

Fio.
J
12-4. Finite unit shearing strain
_ L. -
,
|

H9 denned by tan V'.


I

L
_L

17^
I

The construction of the angle of relative slip illustrated once more in the
1

is
^

auxiliary sketch (Fig. 12-3) to the right of Fig. 12-2 the plane E' a horizontal
is
if

plane.


122 THEORY OF FLOW AND FRACTURE OF SOLIDS

Similarly, in the strained condition of the body the equation of plane E'
may be expressed either by

or in normal form
f
■ti
+ i
yi
+ p
z%
- 1 = 0
(12-18)

a."*' + a/V + o."f -I = 0 (12-19)

where a,",av",a," designate the direction cosines of the normal to plane


E' and I the distance of plane E' from the origin O. By comparison from
Eqs. (12-18) and (12-19):

1 » -L " -1 1 — _L ^ _L ^ 1

(12-20)

For the intercepts x/,yt,zi' according to Eqs. (12-6) and (12-17) we


must also have

Xi' = V\i Xi = V\, O, y/ = V\i - » Zi = VX, -


fl*
; (12-21)

hence

O, = — ,— - > Oy' = — .— > a, = '


VTi r \/X2 r \/\3 t
(12-22)
=
(a,2/Xi) + 5/7x5 + (a.vx,)
The directions a" = a*"2, 6" = a„"2, normal to plane c" = a."2 of the E'
are therefore expressed by means of the directions a,b,c as follows:

-„ o/Xi
'
(a/X,) + (6/X,) + (c/X,)
fr/X,
b" ' ? (12-23 1
(a/X,) + (6/X*,)
6/X,) + (c/X,)
c/X,
c7
(o/X,) + (6/X,) + (c/X,)

and the distance i may be expressed by means of the directions a,b,c or


a",b",c" as follows:

p = = r2(Xia" + x*6" + x,c")


wk)
- <12-24)
(«/x.) + + (c/X3)

The cosine of the angle 4' under which the line OP' = r' = r is \/\
inclined with respect to the distance I measured along the normal to plane
FINITE HOMOGENEOUS STRAIN WITHOUT ROTATION 123

E' is equal to

cos i>
= ir' = — 1-=
r vX
(12-25)

and the unit shear y = tan ^ is therefore given by the equation

72
= tan2 * = — i-7
cos2 ^ -1=^-1 f2

or after replacing I and X through their expressions given above:

t2 = (x,o + \,b + \3c) (r


\Al
+ r + r) ~
A2 A3/
1 - (12-26)

The same expression for y could have been derived by constructing


the tangent plane parallel to plane E' to the strain ellipsoid. Under
reference to Eq. (12-96) expressing the equation of the strain ellipsoid,
the tangent plane in point x',y',z' of this ellipsoid has the equation

Xi X2 X3

Since the coordinates of the points of tangency are x' = -\/Xi aI, • • •

the equation of this plane is also

and since this plane is parallel to the plane E' given by Eqs. (12-18) or
(12-19) the further steps for obtaining the expression for cos 4/ would be
identical to the ones just described.
Equation (12-26), after multiplying out on its right side and simplify
ing, can also be expressed as follows:

7" = X3(Xt
- X2)2a6 + X,(X2 - \3)2bc + X2(X3
- XQ2ca . (12-28)

This and Eq. (12-12) or


X = + X2fr + X3c (12-29)

determine the unit shear y and the elongation X for any direction a,b,c.
To these equations we may add the two conditions
a + b + c = 1 , XtX2Xs = 1 . (12-30)

The advantage of using our new variables X for elongations instead of


the unit strains t becomes apparent, if we compare Eqs. (12-28) to (12-30)
valid for a state of finite strain with the corresponding three equations
124 THEORY OF FLOW AND FRACTURE OF SOLIDS

which were found for a state of stress [Eqs. (10-11), (10-12), and (10-17)].
The comparison discloses that two equations in each group are identical
in form. The expression for a finite shear y [Eq. (12-28)] differs, how
ever, from the corresponding Eq. (10-17) for a shearing stress t in two
important respects: (1) The former one contains in each of the three
terms one factor more (such as X2,Xi,Xj) than the corresponding terms
have for t. (2) The planes of the unit shear y appear tilted relative to the
direction (a',b',c') while the planes in which the shearing stresses t acted
were perpendicular to the directions (a,b,c) to which they referred.
12-2. Graphical Representation of State of Finite Homogeneous
Strain. Thus a graphical representation of such a state of strain becomes
possible in a plane figure in which the elongations X and the unit shears y
are plotted as rectangular coordinates. In an analogous manner, as
described by 0. Mohr, we shall attempt to construct the image of an
octant of a unit sphere
a,2 +V + a,2 = 1 (12-31)

on a X,7 plane. To a point Pfa,.Oy.a,) on this sphere embedded in the


body in its originally unstrained condition corresponds a point Q in a
plane figure with the coordinates X and y. We shall investigate the
properties of the images in this plane of certain circles which we may
choose on the sphere. For this purpose we have to solve the three equa
tions (12-28), (12-29), and (12-30) for the directions a,b,c. This results
in the following expressions:

- X,)(X - +
X(X, - X,)(X. - X,)
x,(x
C, , = X3) y>
a '

- X,)(X - X,) +
.
a"
, X3(X
- X,)(X2 - X.) 72
' (12-32)

X,)(X -
X(XS

- X,)(X, - X,)

, ■
X3(X XQ + 72
c a*
X(X,

There are three pairs of poles such as P\,Pi,P3 on the sphere Eq. (12-31)
through which pass three principal circles or equators a, = 0, a, = 0,
ay = 0. The image points Q\,Qi,Q$ corresponding toPi,P2,P3 lie on the
X axis (since 7 = 0) in the distances Xi, X2 and X3 from the origin. Con
sider, for example, the equator a, = 0. From the last of the three equa
tions (12-32), it follows that when at = 0

X,(X - Xi)(X - X,) + 72


= 0 . (12-33)

This is the equation of an ellipse for the variables X,7. Rewritten in the
standard form, it is
FINITE HOMOGENEOUS STRAIN WITHOUT ROTATION 125

L X,
X. + x2y
x2>

A, - XjV /xt-x,v = 1 (12-34)

\ 2 / \2v/ViX2/
The center C of this ellipse (Fig. 12-6) has the coordinates

Xt + X*
,0;
2

the semiaxes are equal to A =


X*)/2 \/XiXj.
(Xi

X2)/2 and B = (Xi —

Since according to Eq. (12-30) XiXjX3 = 1 (incompressibility condition),


B is also equal to B = (Xi — X2) y/\s/2. There are three such principal
ellipses in the X,7 plane corresponding to the three equators a, = 0, . . .
of the unit sphere. They have the equations

(X
- XQ' "r 2l , _ 1
A* B* (12-35)

Table 12-1 gives the values of the center distances Xc and of the semiaxes

Table 12-1. Principal Elongation Shear Ellipses

Center Semiaxes of ellipse*


Equator distance
X, A B

- X,)/2 - X^/2
a, = 0 (X, +X2)/2
+ X,)/2
(Xi
- X,)/2 VxT
Vx, - Xa)/2
(Xi

\/x^(x, - X.)/2
a, = 0 (X, (X, (x2
av
= 0 (X, + X,)/2 (X.
— X,)/2

• The terms in the columns to be taken with their absolute values.

A and B for the three principal elongation-shear ellipses. The semiaxis


B represents the maximum values of the unit shears or the three principal
unit shears which occur along the three equators. Figure 12-5 shows
a projection of the unit sphere on the plane a, = 0 and Fig. 12-6 the
principal elongation shear ellipse [Eq. (12-34)] for a, = 0. Since on the
equator o, = 0, a, = cos a, a„ = sin a from Eq. (12-7)

X
_ (Xi

X2) cos 2a
2

and from Eq. (12-28) y = \/% (Xi X2) sin 2a/2. The position of -
point Q on the equator a, = 0 is determined through the familiar ellipse
construction from an angle 2a as shown in Fig. 12-6. The same rules
apply for constructing the points Q of the other principal ellipses from
126 THEORY OF FLOW AND FRACTURE OF SOLIDS

Fio. 12-5. Fio. 12-6.


Fios. 12-5 and 12-6. Principal elongation ellipse.

", i SHEAR
UNIT

o3 \
ELONGATION

Fig. 12-7. Principal elongation ellipses.

, UNIT
mi SHEAR

MERIDIAN
CONST
''/a/

03 X
ELONGATION

Fio. 12-8. Graphical representation of elongations X and shears 7.


FINITE HOMOGENEOUS STRAIN WITHOUT ROTATION 127

their auxiliary circles, which are also shown in thin lines in Fig. 12-7.
These three principal elongation shear ellipses thus represent a state of finite
strain in the \,y plane in Fig. 12-7, for which the principal elongations are
Xi,X2,Xj satisfying the condition of incompressibility = 1.

It can be shown in a similar manner that the images of parallel circles,


such as a, = const, . . . and of meridian circles such as Oy/a, = const,
. . which can be traced on the unit sphere [Eq. (12-31)] become ellipses
.

in the X,7 plane. One of their axes coincides always with the X axis and
the other axes are parallel to the y axis. Two such ellipses are traced in
Fig. 12-8 passing through point Q. The images of the parallel circles of
the sphere in a \,y plot do not consist of concentric ellipses.
12-3. Graphical Representation of State of Finite Strains through
Circles. The elongation X and shear y were expressed in Sec. 12-2 by
means of the directions a,b,c. These defined a straight line in the
unstrained condition of the body. It may be of interest to establish
expressions for X and y which depend on the directions a',b',c' defining a
straight line in the strained condition of the body.
From Eqs. (12-9) or

Xo^ b c = — (12-36)
Xi Xj X3

and
a + b + c (12-37)
we obtain
I = ?L + ^ + i. (12-38)
X Xi Xj X3

hence
a
v_ c
X, X2 xi
a = c = (12-39)
a-+b-+c-
Xj
a-+b-+c-Xj a-+b-+c-
Xi X3 Xi X2 Xi X2 A3

After introducing these expressions for a,b,c in Eq. (12-28) we find that

From Eqs. (12-38) and (12-40) an interesting conclusion can be drawn.


Suppose that we define two new strain variables X' and 7' by assuming
that
X'-i, y' =
l (12-41)
128 THEORY OF FLOW AND FRACTURE OF SOLIDS

Equations (12-38) and (12-40) are then transformed in

X' = XiV + X,'6' + X3V (12-42)


- WM - XQW + (X,' - XiQW
,

yi = (x/ + (X,' (12-43)

As a third equation we may add

a' + 6' + c' = 1 (12-44)

A comparison of these last three equations with the corresponding equa


tions for a state of stress [Eqs. (10-11), (10-12), and (10-17)] shows that
both groups are now identical in form. Thus, if the components of stress
in these latter equations
<T T <T\ IT} (Tj

are replaced by the components of strain

X' y' Xt' X2' X,'

Eqs. (12-42) and (12-43) are obtained.


Consequently a state of finite strains (Xi'jXi^Xj') can be represented
by a triple family of circles in a \',y' plane as a state of stress was repre
sented by its Mohr's circles in a <r,r plane. The substitutions in Eqs.
(12-41) have the effect of transforming the plane figure which was

Oy-0

Fio. 12-9. Elongation unit-shear diagram.

described in Sec. 12-2 for the elongation shear ellipses into a figure con
taining only circles. While the abscissas X' in the X',?' plane are the
reciprocal values of the elongations X, the variables y' are given by the
tangents of the angles of the lines OQ drawn from the origin 0 to a point Q
in the \,y plane with the X axis (Fig. 12-8).
An example is shown in Figs. 12-9 and 12-10. In the former one the
three principal elongation shear ellipses are plotted in the X,7 plane. As
FINITE HOMOGENEOUS STRAIN WITHOUT ROTATION 129

an example for the principal elongations the values,

Xi =
I Xj = = 10526 ' Aa =
S = 03800
were assumed (satisfying the condition of incompressibility XiX2A3 =
1).
After introducing the reciprocal elongations
2 1Q
Xi» = = 0.400 A2' = g = 0.950 A,' = ~
^0
= 2.632
\ , ,

the corresponding plane figure is shown in Fig. 12-10, in which A' = 1/A
and y' = y/\ are the coordinates. Figure 12-10 contains only circles.
The images of the meridian and parallel circles, which would be plotted

Fig. 12-10. Reciprocal strain diagram.

on a unit sphere embedded in the strained body would all be represented


by circles in the A', y' plane of Fig. 12-10.

12-4. Strain Components Expressed through Instantaneous Directions in Strained


Body. In Sec. 12-1 the elongation X and the unit shear y were expressed as a function
of a,b,c or the directions of a straight line which was assumed embedded in the body
in its original unstrained condition. It is preferable sometimes and equally important
to know the elongation along a given direction or the unit shear in a given plane which
are chosen in the strained condition of the body, and one wishes to have the expres
sions for the strain components in function of the instantaneous directions embedded
in the strained body. Equation (12-38) expressed the elongation X in the direction
a',b',c' as a function of the latter quantities. We must not overlook the fact, however,
that the unit shear y [e.g., as expressed by Eqs. (12-28) or (12-30)] was obtained for a
certain plane E' in the strained condition of the body which had an oblique orienta
tion relative to the direction o',6',c' embedded in the strained body. What would be
preferred is to have the expression for a unit shear in those planes which are perpen
dicular to the given direction a',b',c' in the strained body.
A simple consideration may furnish us this expression for a unit shear. We desire
130 THEORY OF FLOW AND FRACTURE OF SOLIDS

to obtain an expression of y as a function of the directions a",b",c" of the normal to


plane E'. For a",b",c" we found the values [in Eqs. (12-23)]

a b_
c

Xl »•" — X*
c" x (12-45)

Xi Xi Xi Xi X] Xj Xi X2 X|

We have to solve these expressions for a,b,c, which are not independent, but to which
we must add the condition a + 6 + c = 1. After multiplying a" by Xi, b" by Xi,
and c" by X3 and adding,

X,o" + X^" + X,c"


- + -+-
Xi
r
Xj Xj
(12-46)

From Eq. (12-45) we obtain

Xia" , X26"
Xia" + Xj6" + X,c" Xio" + X,fc" + X,c"

c = ^
Xia
„ ,
+ "\„
Xj6
,. „
+ X,c"
(12-47)

and after inserting these values in Eq. (12-26),

(X,a + X26 + X^ )*

A Becond, equivalent expression for y" may be derived by using Eq. (12-28) instead of
Eq. (12-26), which gives

7
, = (X, — X,)V'6" + (X, — \,Wc" + (X, — X,) W U*«J
(X,a" + X,6" + X^")2

We may now consider a",b",c" as an arbitrary direction embedded in the strained


body. Equations (12-48) or (12-49) express the unit shear 7 in a plane w-hich is
perpendicular to a",b",c". The elongation X" along the same direction [according to
Eq. (12-38)] ia given by

x^xT +
xT+t
12-5. Natural Strain and Natural Shear. In investigations dealing
with finite homogeneous distortions in such elastic substances as rubber
or with the flow of the ductile metals in the industrial forming processes
it is frequently advantageous to make use of the components of the
natural strain and shear.
The natural strain was introduced in Chap. 8 when the stress-strain
curve of the tensile test was discussed. To emphasize its significance
once more when describing general states of finite homogeneous strain
a natural strain i in the direction OP' = r' according to P. Ludwik and
A. Leon is defined by the increment dl = dr'/r' equal to the change of
the length dr' divided by the instantaneous length r' of a straight line
FINITE HOMOGENEOUS STRAIN WITHOUT ROTATION 131

embedded in the material. Since r'2 = r2X, 2r' dr' = r2 d\ where r


denotes the original length connecting the same material points in the
unstrained condition of the body:

dS
di (12-51)
r> 2X
and the natural strain:

« "
£ y - In j = In (1 + t) = i In X , X = e2' . (12-52)

For incompressible material:


X,X,X, = e*»'+',+'.> = 1 (12-53)
and after taking logarithms, for
the natural principal strains
ii + h + U = 0 . (12-54)
Suppose that OA, OB, and OC
in Fig. 12-11 are three mutually
perpendicular directions of unit
length in the strained condition
of the body (in Fig. 12-11 the
directions x,y,z are the principal
directions of strain and ii,ii,i3 the
principal natural strains). Let
the principal directions remain the
same when the strains increase by
the increments dii,dii,dii- Ac
cording to Eq. (12-54)
dii + di2 + dh = 0 . (12-55) Fig. 12-11. Increments of natural strains.
The points A,B,C are displaced
in new positions A',B',C Then let AA' = dsa, BB' = dsb, CC = dsc
(Fig. 12-11) and denote the components of these three infinitesimal dis
placement vectors in the directions OA, OB, and OC, respectively, by

d«o , . . . , dia , dyab , dyac ,

dsb , . . . , dyba , dh , dykc , (12-56)


dsc , . . . , dyca , dyA , die ,

The six quantities dy^dy^dybc, . . . express the small angles by which


the material lines which coincided with OA, OB, and OC are tilted with
respect to the planes of these three lines when the strain increases by an
increment.
132 THEORY OF FLOW AND FRACTURE OF SOLIDS

Three increments of the natural shears are herewith defined, namely,

dya = dyu + diet , \


dyb = dye* + dyae , I (12-57)
dye = djab + dyta . j
The equations (12-57) may serve to define the natural shears:

76 = 7ca + 7« , (12-58)
|
7r = 7o* + 7fca ■ )

It is understood that this integration is to be carried out under the


assumption that the three directions OA ,OB,OC remain fixed in space and
unchanged in their position relative to the principal directions x,y,z of
the strain. As a consequence, while the strain increases, new material
lines must pass through these three fixed directions. It must therefore
be noted that the natural shears defined by Eqs. (12-58) no longer refer
to the same material lines in the deformed body when the distortion
increases. They refer to those material lines which successively sweep
through a given triple of mutually perpendicular directions held in fixed
position in space.
A few examples treated in Chap. 14 will illustrate their usefulness.
CHAPTER 13

FINITE PLANE STRAIN. EXPRESSIONS DEVELOPED


IN TERMS OF THE NATURAL STRAINS
When one of the principal strains, e.g., t3, vanishes, we obtain the case
of plane strain.
13-1. Pure Shear. A homogeneous finite plane strain is called a pure
shear if the body is uniformly strained in two given directions which are
perpendicular to each other and if
the volume in the strained condi
tion remains constant. Choosing
these two directions for the x and
y axes, the equations for a pure
shear are obtained from the general
expressions which were developed
in the preceding chapter by assum
ing that the principal strain e3 = 0.
It may be just as simple to derive
them directly.
If «i and c2 denote the strains in
the principal directions x and y,
respectively, the coordinates x',y',z' Fio. 13-1. Pure shear.
of a point P' in the strained con
dition of a body distorted by a pure shear are expressed by
= - =
ar*

e,)x ti)y
y1

+
+

z'
(1

(13-1)
(1

z
,

x,y,z being the coordinates of the same material point in its original
position P.
Furthermore, for incompressible material,

+ «,)(! + «,) =
(1

(13-2)
!

Using again the elongation in the direction OP' (Fig. 13-1) in which
X

the strain
is
t,

= +
(1

(13-3)
X

«)2

and Xi = a)1, Xj = principal elongations,


+

«j)2 as the the con


+
(1

(1

dition of incompressibility [Eq. (13-2)] also expressed by


is

133
134 THEORY OF FLOW AND FRACTURE OF SOLIDS

XAs = 1 (13-1)
and
x' = \/\[ x ,

1 (13-la)
y1 y ,

assuming that the point P(x,y) lies in the plane z = 0. The length of
a line OP = r which is inclined under the angle a with respect to the
x axis is changed in a length OP' = r' inclined under an angle a'. From
Eqs. (13-la),
r' cos a' = \/X~i r cos a ,

1 (13-16)
r sin a

The elongation X along 0/" is

r _ x' + J/
Xi cos2 a + r- sin2 a
Xi
or

cos 2a (13-5)

and from Eqs. (13-16)


tan a = Xi tan a' (13-6)

In order to express the unit shear y consider the right angle at point
P(x,y) between line OP and the straight line E normal to the direction
OP = r in Fig. 13-1 drawn through P. After strain E will take a position
E'. HE is inclined under an angle 0 with respect to the positive x axis
and E' under the angle 0', since E and E' are also corresponding direc
tions in the unstrained and in the strained condition of the body, respec
tively, we may substitute in Eq. (13-6) the angles /3 and 0' instead of the
angles a and a' and have also:

tan = Xi tan
/?'

/3 (13-7)

But = (ir/2) a and =


(t/2) + a" where a" denotes the angle
+

0'
/3

of inclination of the normal OQ' to line E'. Therefore

tan a" = Xi tan a


(13-8)
.

Suppose that the two straight lines and E' represent the traces of
E

two planes which are perpendicular to the x,y plane and that the angle
a' — a" =i^. the angle between OP' and the normal 0Q' to plane
is
f
FINITE PLANE STRAIN 135

E' (Fig. 13-1). The unit shear y in the plane E' is equal to

tan a' — tan a"


y = tan = tan — =
<p (a' a") 1 + tan a' tan a"
(13-9)

With the values of tan a' and tan a" taken from the Eqs. (13-6) and
(13-8) we find the unit shear y equal to

sin 2a
Kxi~^)si (13-10)
.

From Eqs. (13-5) and (13-10) after squaring and adding

+ 7J J-]S (13-11)

in the variables X and y the equation of the strain circle for a finite pure
shear is obtained.1

ELONGATION

Fig. 13-2. Pure shear.

The expression for the unit shear y [Eq. (13-10)] could also have been
obtained directly from another form of the condition of incompressibility.
Consider a unit square in the unstrained condition of the body with the
sides oriented under the angles a and (?r/2) + a and let the corresponding
elongations be X„ and X„ in the strained body. The square must be dis-
1 We note that the largest of the three principal elongation ellipses in the graphical
representation of a state of finite strain according to the rules laid down in Sec. 12-2
must become a circle for a state of pure shear. Furthermore, upon inspection of the
values of the semiaxes of the principal elongation ellipses given in Table 12-1, p. 125,
we see that for a pure shear the two smaller ellipses become similar and the

B, - By = A,AV .

An example is shown in Fig. 13-2 for Xi = 2, X2 = }4, X3 = 1. The strain circle has
the equation (X — 5£)2 + y2 =
136 THEORY OF FLOW AND FRACTURE OF SOLIDS

torted in a rhomboid of equal area. The latter condition is expressed by


the equation

X,X„ - -\-:
cos2 4*
= 1+7* (13-12)

which permits the computation of the shear y.


The strain components may be expressed finally as a function of the
directions under strain :

a" = cos2 a" , b" = sin2 a" , c" = 0

using Eqs. (12-49) and (12-50). We wish to know the elongation X"
along the normal OQ' to the plane E' in the direction a" in the strained
body (instead of X along the line OP' in Fig. 13-1). This and the unit
shear y in the plane E' perpendicular to the direction a" are given by

y =
(X!
- X,) Va"b"
- (l/X,)] 2a"
x,a" + x26"

Xi
[Xi
+ (l/X,) + - (l/X,)] [Xi
sin
cos 2a'
(13-13)

13-2. Pure Shear Expressed by Natural Strains. The preceding expressions for
the components of strain may be presented in more symmetric form by introducing
instead of the strain components X and y the natural strains.
Let the natural strain1 i in the direction in which X was expressed be defined by

X = (1 + .)« = e«! . (13-14)

For the principal values for a pure shear we have Xi = 1, 5« = 0 and because
XiXi = 1, it = — Si, also

\ ±-\ = cosh 2i, - ±-\ - sinh 2i,


i

(13-1 4a)
+

(Xi
(xt

.
,

Consider a pure shear = — J0, <j ~ «o, for which the dimensions of the body are
ii

shortened in the x and lengthened in the direction, assuming that the constant <o or
y

"the intensity measuring the pure shear" e0 > 0. we make use of the abbreviations
If

= sin 2a = cos 2a
- sin 2a"
s

c
,

s" c" = cos 2a"


,

= sinh 2i0 = cosh 2l0


C
S

(13-15)
,

The natural strain


in the direction OP' = in accordance with what was said in
1

r'
i

Sec. 12-5 defined by the increment di = dr' /r' equal to the change in length dV
is

divided by the length r'. Since r" = Xr1, 2di = dX/\. The latter expression inte
grated gives 2i = In or Eq. (13-14).
X,
FINITE PLANE STRAIN 137

the expressions for the elongations X and X" may be rewritten as follows:

X = e»« - C - Sc , X" =
C + Sc"
1
(13-16)

and those for the unit shear:

y = Ss =
C -
Ss"
Sc" (13-17)

to which we may add the two angular transformations

tan a = e-s!° tan a' , tan a" — e~"° tan a (13-18)

connecting the three angles a, a', a".1


A certain rhombus A BCD (Fig. 13-3) is dis
torted into a congruent rhombus A'B'C'D' in
which the acute and the obtuse angles are
interchanged. The angle a = a0 under which
the side AB of the first rhombus is inclined
with respect to the x axis has a tangent which
must be equal to tan a0 = 1/(1 + e») = e~«°.
When the side AB is displaced in the position
A'B' it retains its length; therefore X = 1,
(J = 0), and the angle ao for the "unstretched"
rhombus must satisfy the condition [see P^q.
(13-16)] 1 = C - Scot
C - 1
Fig. 13-3. Unstretched rhombus.
cos 2a0 tanh to . (13-19)

Let us consider now two sets of perpendicular directions, either in the unstrained, or
in the strained body. For the first ones we choose the directions v and n of the two
lines OP and OQ in the unstrained body for which a, = a and aM = (ir/2) + a (Fig.
13-4). If the square OPRQ under the strain <0 has been distorted into the rhomboid
OP'R'Q' (Fig. 13-4) suppose that the elongations along OP' and OQ' are equal to X,
and Xp. Then according to Eq. (13-16)

X, = C - Sc, X„
= C +Sc and K + X„
= 2C = const (13-20)

If one desires to express s, c by


1
c',

or by s", c" or vice versa, by trigonometry


s',

from Eqs. (13-18) follows that


it

+ Cc' Cc -s
C S

+ Sc' +Sc' -Sc -Sc' (13-18a)


C

Similarly,

+Cc -
r - Sc" Cc"
- Sc"
S
C S

(13-186)
'

+ Sc Sc
'
+
C

C
C

Equations (13-185) are obtained from Eqs. (13-18a) after replacing a by a" and a'
by a.
138 THEORY OF FLOW AND FRACTURE OF SOLIDS

and we see that the sum of the elongations resulting from a pure shear for two direc
tions r and p, which originally in the unstrained condition of the body were perpendicular,
is a constant = Xi + X2 = 2C.1

Fig. 13-4. Pure shear.

For the second set of perpendicular directions n and m assumed in the strained
body, we choose the direction of the normal n to a "section" P'R' (or to OQ') and the
section (OQ1) itself m. Under the intensity of shear ia the normal n makes an angle
an" = a" and the section OQ' coinciding with the direction m an angle

am
_
a = * 1 11
+ <*
2

with the x axis. Therefore, from Eq. (13-16),

1
C + Sc" C - Sc" and -L
An
+ JL
Am
= 2C = const (13-21)

Equation (13-20) expresses a well-known property of an ellipse. A circle of


1 a

radius equal to unity around the origin 0 is distorted into the strain ellipse

x"

Xi
11"
+y~
x,
- 1.

For a pure shear Xi = 1/Xi = e'*<>,the elongations X, and (which are equal to the
X,,

squares of two oblique radii of the strain ellipse) are measured along two conjugate
diameters. But the sum of the squares of conjugate radii in an ellipse equal to the
is

sum of the squares of the major and minor semiaxes, viz.,

X, = f5?« e~"» = cosh 2J0 = 2C


+
+
X„

which Eq. (13-20).


is
FINITE PLANE STRAIN 139

The sum of the reciprocal values of the elongations for two perpendicular directions
in the strained body is constant. Evidently X„" is the same elongation as \, and
X_" = X,.
The unit shear y according to Fig. 13-4 is composed of two finite shears yr„ and y„r.1
The shear y is obtained from y,? and •»,. through applying the addition theorem of the
tangent function:

7'M + T>'
(13-22)
1 — 7»^7m»

for the tangents of the corresponding angular differences.


The question may now be raised, whether the standard conventional shearing strains
which were defined by the tangent functions of the angular differences such as

7 = tan (a' — a") , yrll = tan (a' — a) , y,,, = tan (a —


a") (13-23

represent the most suitable means for expressing angular distortions under large
.strainsand whether other measures for them could be proposed in analogy to the
procedure which led us above to introduce the natural strains i instead of the con
ventional ones. This will be considered in the following section in which new expres
sions for the natural unit shears are proposed.

13-3. Work Done under Finite Pure Shear. The Natural Shearing
Strains f
and y. A unit cube of material in a state of pure shear is under
the action of a compression stress
<ti
= —To and a tensile stress

<7j = to(t0 > 0)

in the directions of the x and y axes.


When the intensity of stress t0 in
creases by a small amount, the sides
of the cube contract by 8ii = —8i0
and lengthen by Sit = SiB, respectively,
and the increment of the work done is

SW = <7i Sti + <72dit = 2t0 Sio . (13-24) Fio. 13-6. Pure shear.

Or consider a prism of material (Fig. 13-5) when the strain has reached
the values «i = —to and tj = «o- To produce this strain a state of stress

a = —to cos 2a" , I

t = to sin 2a "
(13-25)
J

is required, where a" designates the angle of the normal to the oblique
face of length c of the prism, a the normal, and t the shearing stress in it
(Fig. 13-5). When the strain i0 increases by an increment 5«0 only one
1
The arcs in Pig. 13-4 denoted by ytli and y,,, are symbolically meant. They
represent in fact the angles the tangents of which are equal to these unit shears.
140 THEORY OF FLOW AND FRACTURE OF SOLIDS

of the four resultant forces of stress a\b, am, <rc, and tc, namely, re does
work. For a pure shear
ha 56
6(0 =
— —
a
= -r
b

and the center point C" of the oblique face c of the prism moves by the
amounts &a/2, Sb/2 in the x and y directions. The triangle CDC"
showing these displacements of point C" is similar to the triangle a,b,c,
and point C" moves just along the face c. Since C'C" = c 6i0/2 the
increment of work per unit of volume of the prism must be equal to

^-°
SW = cr = 2r0

J;
5e„ (13-26)

which the same value as that given by Eq. (13-24). (The center point
is

C of an oblique section describes as all other points during the deforma


tion of pure shear an equilateral hyperbola as the streamline of the flow,
a

which for the special point C" also the envelope of all oblique lines in
is

c
their consecutive positions as the strain increases.)
Although either of these simple means furnishes us the expression for
the increment of work SW let us now consider in a more general manner
the increment of work done by the components of stress which act in the
faces of a unit cube oriented in an arbitrary direction in the body when
the strain reaches the value e0. cross section of this cube indicated
A

is
in Fig. 13-6 by the square OP'R'Q' in which the side OP' inclined under

is
an angle a" with respect to the x axis. This square distorted into
is

a
rhomboid OP"R"Q" and unit circle scribed into into an ellipse, after
it
a

the strain increases by the increment 8i0. P'P" and Q'Q" are the arc
elements of the streamlines of flow passing through P' and Q'. Two
small rectangular triangles are indicated in the figures, the short legs of
which represent the small displacements of the points P' and Q' parallel
to the axes. Since, for P'(x',y'),
6x' Sy'
x.
x'
y'

and similar condition holds for point these two triangles must be
Q'
a

similar to the respective triangles in which x', y', and are the sides.
1

Suppose that the directions OP' and OQ' in Fig. 13-6 are the directions n
and m which were introduced in Sec. 13-2 (see Fig. 13-4).
It sensible now to introduce two increments of shear 5f„m and 5y*n
is

(see Fig. 13-6), which designate the small angles by which material lines
coinciding with the two perpendicular directions n and m in the strained
body rotate when the strain has increased by 5i0. They express the
instantaneous change a right angle suffers in strained condition the body
of
a
FINITE PLANE STRAIN 141

and are introduced in a manner analogous to the way in which the incre
ments of the natural strains o*„ and Sim were denned (namely, as the
in the strained body suffers, in the directions n and m).
changes a unit length
By projecting P'P" and Q'Q" on the directions n and m, one verifies that
Sin = -Sim= -
6i0 cos 2a" , (13-27)
87.- = Symn = Sio sin 2a" (13-28)
assuming that the positive directions for these two unit shear increments
are the ones indicated in the figure.

Fig. 13-6. Increment of pure shear.

The increment of work is


bW = <r„ 5«„ + <jm 5i„, + r„m 3-y„m + rm„ by (13-29)
When
<r = cn = —to cos 2a" , r = rnm = t0 sin 2a

= — Tm„ =
= To cos 2a" -T0 sin 2a"
~2 +
a ' Cm .

' N
142 THEORY OF FLOW AND FRACTURE OF SOLIDS

The increment of work becomes equal to

SW = 2t0 5io(cos2 2a" + sin2 2a") = 2r0 8i0 (13-30)

in agreement with the value which was formerly obtained. The portion
2t0 5«o cos2 2a" represents the work done by the normal stresses aH and
<r„ and 2t0 5«o sin2 2a" the work of the shearing stresses.

The angular change of the right angle between the directions n and m
composed of the sum

Sy = Bynm + 8ymn = 2 5«„ sin 2a" (13-31)

is defined as the increment of the natural shearing strain Sy. Since

2 5?o
= — (Sii — 8«j) ,
also
Sy = - (Sii - 8h) sin 2a" . (13-32)

Si„ — Sim and Sy satisfy the equation of a circle:

(5i„
- Simy + Sy* = 4&o2 . (13-33)

We may now compute a finite expression for y by integrating Eq.


(13-31) from the unstrained condition of the body to a strained one. But
we have to make up our mind precisely how this integration from one
strained position to another one should be carried out.
Suppose that we maintain during the integration the angle a" = const.
Then, from Eq. (13-31),

y = 2in sin 2a" = - - t,)


(e, sin 2a". (13-34)

This natural shear of the first kind y is proportional to sin 2a" similarly
as the shearing stress r according to Eq. (13-25) is proportional to sin 2a".1
On the other hand, we may follow the changes in position of material
lines which coincide instantaneously with the direction OP' of the normal
n of a given material section P'R' or OQ' in Fig. 13-6. When the intensity
of the strain i0 increases P'R' or OQ' change their positions in space.
Each new position of these two parallel lines refers to the same material
section in the body. The normal n to this section (considered as a geo
metric line) rotates by the same small angle 8a" by which the section
1 We may visualize this shear of the first kind f. It consists of the sum of the angular
increments which were swept through the sides of a right angle held in a fixed position in
space during the distortion of the body. 7 is the sum of the infinitesimal angles by which
consecutive lines of the body move past the edges of the right angle. (See sketch in
Fig. 13-7. This right angle could be traced on a glass plate placed above point 0 in
the body and held in a fixed position in space when the body distorts.)
FINITE PLANE STRAIN 143

rotates. We may easily compute 8a" from the second of the two trans
formation equations (13-18):
tan a" = e-2* tan a (13-35)

by differentiating it with respect to i0 keeping the angle a constant. This


gives

:— r, = — 2e-2,° Si0 tan a

or in conjunction with Eq. (13-35) :

5a" = -5«„sin2a". (13-36)

We note now that the increment of shear which was introduced by


Eq. (13-31) is also equal to
hy = -2 5a". (13-37)

Integrating Eq. (13-37) from the unstrained condition of the body in

tttt\

Fig. 13-7. Increments of natural shear by = &ynm + *7i»ii.

which a" = a until the intensity of strain has reached the value i„ we
obtain now
7 = 2(« - a") (13-38)

as the equation defining natural shearing strain of the second kind y.


the
During this last integration the angle a was maintained constant.1
1 When a material line OP'
turns around O in a new position OP" (Fig. 13-6) by a
small angle ilnm when i0 increases by St0, the geometric'line, which coincided with it
first and was normal to the material section P'R', turns in the opposite sense by the
same angle ( — 6a") to become again the normal to the displaced section. Thus a
material line such as OP" becomes separated by twice the angle of rotation of the
material section (to which it was originally perpendicular) from its normal for each
increment of strain St0.
144 THEORY OF FLOW AND FRACTURE OF SOLIDS

Neither of the two kinds of natural shears y and y refers to the same
material lines v and n which originally formed a right angle. The
tangent of the change of the angle of inclination of these originally
perpendicular material lines under strain was defined as the conventional
unit shear y — tan ^ [Eq. (13-10)]. The natural shear y is the integral
of the angular increments swept through both sides of a right angle held
in a fixed position in space while the natural shear y is the integral of the
angular increments swept through the geometric normal of a given material
section during the process of straining. The advantage of using y or y
becomes evident when considering work hardening, or plastic strains of
large magnitude, or the work done by the shearing stresses, etc.

In preceding considerations we were following trends of thoughts which have their


analogy in two reference systems used in hydrodynamics.1 The Euler system describes
what happens during the motion of a fluid at a given geometric point fixed in space.
To this corresponds the manner in which the natural shear of the first kind was
defined as far as the measurement of shearing strains is concerned. In the second,
Lagrange system one desires to determine the actual orbits of the particles of the fluid,
their velocities along their paths at given times, etc. The use of the conventional
unit shears y, which defined the change in the angnlar position of two originally perpen
dicular material lines or sections in the body, corresponds to this. Finally the natural
shear of the second kind may be termed as a kind of semi-Lagrange variable. Thus
in the case of -f new particle lines pass across one geometric line (the normal to the
Bection), while in the case of f they pass across two geometric lines (both fixed in
space).
It may be of interest to note that the maximum of the natural shear -f = 2 (a — a")
under some given strain t„ occurs along an angle a" and a corresponding angle a which
define the directions of the normals of the sides of the unstretched rhombus (Fig. 13-8).
To determine the maximum value of as we consider various directions a between
the values zero and t/2 at a given strain i0 we have to compute the angle a" from the
relation
tan a" = e_,.° tan a
which is
a" = tan-1 (e_,.o tan a)

and to differentiate the expression for -f with respect to a while e0 = const. From

y = 2(« — a") = 2a — 2 tan"1 (er*U tan a)


we find

At the maximum (d-f/da = 0) the bracket must become zero:

C — 1
c = cos 2a = g— -

1 Prandtt-, L., and O. G. Tietjens, "Fundamentals of Hydro- and Aeromechanics,"


Vol. 1, p. 69, McGraw-Hill Book Company, Inc., New York, 1934.
FINITE PLANE STRAIN 145

Therefore

IS + C -
un.-VRR-V 5 -C +
1
= *'.
1

and
tan . e-,,° tan a

But these two angles belong just to the normals of the sides of the unslretched rhombus in
the first quadrant before and after strain was applied (Fig. 13-8). We can see that
while the maximum value of the natural shear f
occurs at an angle a" = 45 deg in a
fixed direction in space, the maximum of -f shifts its position nearer to the axis of the
compression stress <n = — t0 as the strain <0 increases (Fig. 13-9).

tftux.

Fio. 13-8. Unst retched rhombus. Fig. 13-9. Values of 1 = /(a").

With the values of tan a and tan a" which were just computed, one finds for the
maximum value of -f the equation

e** — e~^
tan
(¥)- tan (a —
a") = sinh (i,

determining 7m„ at a given strain i0.


Example : Suppose that a perfectly elastic material is deformed
by a finite pure shear.
For an incompressible and isotropic material we should assume that only two
elastic
elastic constants are required, the modulus of elasticity E and the modulus of shear G;
furthermore, G = E/3. From the familiar stress-strain relations expressing Hooke's
law for infinitesimal strains and for an incompressible material we should obtain for
the increments of the natural principal strains the equations

T'
5<Tl
E «ii = AVi E H-. Sat —

For a pure shear,


&<ri = —iai = — 4t0 , Sii = — Sit = —Sen

so that T0 and to are the measures of the intensity


of stress and of strain. These equa
tions may be integrated at once since they refer to the same material lines during the
integration. Therefore, for a finite pure shear,

?1 =
— ?2 = —to = — ■
Kq (13-39a)
146 THEORY OF FLOW AND FRACTURE OF SOLIDS

For a perfectly elastic material, furthermore,

frj = iy = — .

If we wish to integrate these two equations we have the choice of integrating them
either when the angle a" = const or by assuming that a = const. This remark must
also refer to the increments of the shearing stress t. To distinguish these two cases,
let us designate by if the increment of f when a" = const. Since f = t0 sin 2a",
evidently
if = Jt0 sin 2a" .

Similarly let 4f be the increment of f when a = const. 8f is the "local," *r the "total"
increment of the shearing stress t:
if = 4(t0 sin 2a") .

Hence, after integrating the equation for the increments of shear, we obtain either

1 = 2«0 sin 2a" =


G
(13-396)

V - 2(« - a") - - •

We see, therefore, that the stress-strain relation for the finite natural shears 7 and y
are preserved in their linear form valid for infinitesimal strains.
If we attempted to express the relation connecting a shearing stress t with the con
ventional shear y we should find a fairly complex behavior.

13-4. Simple Shear. If x,y are the coordinates of a point P in the


unstrained and x',y' in the strained condition and r,a and r',a' the cor-

&
/[M.
Flu. 13-10. Simple shear.

responding polar coordinates, a simple shear is denned by the linear


transformation
x' = x + 7.2/ = Kcos « + 7. sin a)
j (13-401
y' = y = r sin a

where the quantity 7, = tan S (Fig. 13-10) is a measure of the simple


FINITE PLANE STRAIN 147

shear equal to the tangent of the angle of tilting 5 of the lines x = const.
The elongation X in the direction a' is

X = j— = cosJ a + 27, cos a sin a + y.2 sin2 a + sin2 a

or

X = 1 + ~
'"
- ~ cos 2a + 7. sin 2a
2
. (13-41)

Let XM and X„ be the elongations corresponding to the angles a, = a and


a„ = a + (ir/2). For any two originally perpendicular directions,
evidently
+ = 2 + 7.s
Xm Xn • (13-42)

A unit square in the unstrained condition of the body with the sides
inclined under the angles a and a + (ir/2) is distorted into a rhomboid
of equal area having the elongations Xm and X„. We have,, for the rhom
boid of equal area,
XmX„ = 1 + t2 (13-43)

which serves for computing the value of the unit shear 7 using Eq. (13-41) :

2
-v
7 = -y sin 2a + y, cos 2« . (13-44)

From Eqs. (13-41) and (13-44) after squaring and adding,

- -
^J +
= 7.2 +
(l ^)
1 72 (13-45)
[x
in the variables X,7 the equation of the strain circle and after letting
7 = 0 the principal elongations are obtained:
y)2
(-^1

=
yjl

+ =
+
^
+

± 7.
^
±
^

1 (13-46)

X„X2

From the last equation we see that «i — a = y, for simple shear.


a

For an incompressible material state of plane strain essentially a


is
a

pure shear combined with rotation of the principal directions of strain.


a

A simple shear 7, therefore also pure shear combined with certain


is

rotation. The principal directions under simple shear satisfy the


a

condition

tan 2« = -- (13-47)
7.
obtained after letting =
in Eq. (13-44). The two angles a satisfying
0
7

Eq. (13-47) define the directions of those two perpendicular lines in the
unstrained condition of the body, which, after the strain applied, will
is
148 THEORY OF FLOW AND FRACTURE OF SOLIDS

become the principal directions of strain. The corresponding angles a'


in the strained condition of the body satisfy the condition

tan 2a = —
(13-48)
7.
X and 7 may be expressed by referring them to the angle a' in the strained
condition (instead of to a) by

i = + - cos 2a' - 7. sin 2a'

Zj-'

^
1

,
(13-49)
5- sin 2a' 7. cos 2a'
X 1

+
The principal directions under a simple shear, and those perpendicular
directions in the unstrained body which after strain become the principal

Fio. 13-11. Principal directions for simple shear.

axes, are very simply found by constructing the " unstretched rhombuses "
in the strained and unstrained body.1

For a simple shear make 7. = tan in Fig. 13-10. OC = and EC = CE' = y,/2.
1

OE, OE' define the directions and lengths y/\ + (y.'/4) of the sides in the two con
gruent rhombuses OEDA and OE'D'A' which represent those lines in the unstrained
and in the strained body which do not change their lengths. (The former ones are
dashed, the latter ones represented by full lines.) The diagonals OD' and OF' define
the principal directions of strain, the diagonals OD and OF the corresponding direc
tions in the unstrained body. m and on' denote the halves of the obtuse and of the
If

acute angles in the two rhombuses, from Fig. 13-10,

OC OC
1

tan ai = tan on' =


,

CD VI — CD'
'

+ (7.74) (7./2)
'
Vl (7.V4) + (y./2)
+

We note that the angle of rotation o>of the principal directions given by
is

tan <o tan (ai —


ai) = = — —
-
2

See also Fig. 13-11.


FINITE PLANE STRAIN 149

In isotropic material a given intensity of a finite simple shear 7, must


be equivalent to a certain magnitude of a pure shear, the intensity of
which may be measured, for example, by the natural strain «0 which was
introduced in Sec. 13-2 and through which the principal elongations for a
pure shear were defined by

Xi = e"° , X, = er** . (13-50)

The equivalent simple shear 7, must have the same principal elongations.
But, from Eqs. (13-46),

g = h+± -!,«'» + «-» - - cosh2;o - !


!

and

= sinh €0 (13-51)
^

defines the equivalent intensity of the simple shear 7,. Since the maximum
natural shear y^ for a pure shear was found equal to 2*0, also

7, = 2 sinh
^ •

These relations define equivalent simple and pure shears for a ductile
metal that has been cold worked by these strains. We may state that a
metal has been work hardened to the same degree by simple or by pure
shear if 7, = 2 sinh i0 and may add that Eq. (13-51) expresses the con
dition that the metal has suffered the same angular distortion, which
may be visualized by the congruence of the "unstretched rhombuses"
corresponding to both strains. It will be shown later for an ideally
plastic substance that the mechanical work required to produce these
two equivalent shears is not the same unless an additional condition is
prescribed, namely, that the plastic substance is deformed under a simple
shear by a system of stresses whose directions of principal stress coincide
continously during the straining with those of principal strain.

Example: Simple finite shear in elastic material. We suppose for an isotropic


elastic material that the condition which was just mentioned must be prescribed and
note that a system of shearing stresses tiv is not sufficient to produce a simple finite
shear 7, because the principal directions of strain must satisfy Eq. (13-47) while the
principal directions of stress of such a state of stress are inclined at the constant angles
of ±45 deg with respect to the axis. The principal directions of strain n and <j of a
simple shear will coincide with the principal directions of stress if also normal stresses
ax and «■„of a certain magnitude are present in the body Resides the shearing stresses
150 THEORY OF FLOW AND FRACTURE OF SOLIDS

T,v (Fig- 13-12). Since the sum of the principal stresses must vanish for a pure shear
for the new state of plane stress, <rx + av =0 and a, = —a,. The principal directions
of stress and strain will coincide if

2rJy T,y 2
<T,
(13-52)
<Jv o* 7*

Since the principal stresses must also satisfy the


equation

<7l = ~<rt - — Vo,2 + T,,S = —TO (13-53)

:. c we have for the determination of a, and r,y the two


\ conditions

—* = —
+ t„2 = to' and (13-54)
Fio. 13-12. Plane stress re
quired to produce finite simple
shear. from which, after using the value for t0 = 2Gi„
from Sec. 13-3, Eq. (13-39a) and the equivalence

condition Eq. according to which i0 = sinh-1 we see that a simple


(13-51)

shear y, in elastic isotropic material requires the stresses

To7„
<r, = — ay
T" - (13-55)
2 Vl + (7.V4) VI + (7.74)
where


to = 2G?„ = 2G In +
VT+^'] (13-56)

These equations are also equivalent to

2G;0
d = — <r, = 2Gi° tanh Jo , t,„
cosh ?o
(13-57)
CHAPTER 14

VECTOR GEOMETRY OF STRESS AND STRAIN.


LINEAR VECTOR FUNCTIONS. TENSORS. VECTOR FIELDS

The concepts of stress and strain were established by Cauchy around


1822. Together with the theory of the potential, of the functions of a
complex variable, the energy principle, and the calculus of variations,
they constitute the foundations on which the mathematical theory of
elasticity and classical hydrome
chanics were developed during the
nineteenth century by Navier,
Poisson, Green, Stokes, Kirchhoff,
Helmholtz, St. Venant, Boussinesq,
Maxwell, Lords Kelvin and Ray-
leigh, Love, Lamb, and others. In
1882, Otto Mohr published his first
paper on the graphical representa
tion of the state of stress. Mohr
remarked that his graphical method
was also applicable to the represen
tation of the distribution of the
moments of inertia in rigid bodies.
In mechanics and physics cases
are frequently encountered in which
the three components of a vector in
space are linear homogeneous func
tions of the three components of
another variable position vector.
We shall devote some attention to
Fig. 14-1. J. Willard Gibbs. (Acknowl
such cases in this chapter, pointing edgements are due to Muriel Rukeyser,
" Willard Gibbt."
to the fact that the state of stress, author of monograph,
Doubleday, Doran & Company, Inc., New
of finite homogeneous strain, and York, 1942.)
of the velocity distribution of the
strains around a point in a deforming material are examples of such
functions. They may therefore be studied from a common point of
view by putting the form of the relationships connecting the mechanical
variables in the foreground. This was admirably achieved in Josiah
151
152 THEORY OF FLOW AND FRACTURE OF SOLIDS
" "
Willard Gibbs's Vectoranalysis around 1881. In it was shown that
these and other related concepts in physics have a common geometry:
they are examples of linear vector functions. The properties of these
functions will be reviewed in the following sections.1
14-1. Linear Vector Functions. A vector is a quantity having magni
tude (an absolute value which is denoted by an italic letter) and direction
and a sense along this direction. Constant vectors may be denoted by
the bold-face letters a,b,c, . . . , variable vectors, for example, by r
or v. Examples: the velocity, the acceleration of a material point in
motion, a force, a moment. A variable vector, for example, may serve
to represent the position of a movable point P in space. If we choose
the origin O of a right-handed rectilinear system of cartesian coordinates
as a fixed point in space, we can represent the position of P by a position
vector OP = r. In accordance with adopted practice in vector analysis
we define position vector r by the geometric sum

r = ix + jy + kz (14-1)

in which i,j,k are the unit vectors and x,y,z the components of r along the
positive directions of the coordinate axes.2
The simplest linear vector function is the product of a scalar (a quan
tity having no direction but only a magnitude) and a variable position
vector r:
r' = cr . (14-2)

The linear function r' of r is here a vector. But such functions may
become scalar quantities or quantities of higher order than a vector in
space.

1 It is not presumed that the content of this chapter will assist the reader in his

attempts to develop special solutions in the theory of plasticity. The less experienced
reader may as well entirely omit studying it. It is hoped the concise tools for present
ing the properties of linear vector functions may prove helpful for further clarifying
some of the physical laws under which materials deform.
2 It is assumed that the reader is a little familiar with the elements of vector analysis

and with the definitions of the scalar and vector product rules. The text above follows
parts of courses on mechanics of continua which were formerly given by L. Prandtl
and by the author at the University of Gottingen based on the classic book by J. W.
Gibbs and E. B. Wilson, "Vectoranalysis," 2d ed., Yale University Press, New Haven,
1901. See also Prandtl, L., and O. G. Tietjens, "Fundamentals of Hydro- and Aero
mechanics," Chap. 6, McGraw-Hill Book Company, Inc., New York, 1934; Durand,
W. F., "Aerodynamic Theory," Vol. Ill (Div. G, chapter on Viscous Fluids by L.
Prandtl, p. 42), Verlag Julius Springer, Berlin, 1935; Geiringer, Hilda, "Geometrical
Foundations of Mechanics," Brown University, Providence, R.I., summer session,
1942, 154 pp.; Brand, Louis, "Vector and Tensor Analysis," John Wiley & Sons, Inc.,
New York, 1947, 439 pp.
VECTOR GEOMETRY OF STRESS AND STRAIN 153

Let r variable position vector OP drawn from a fixed point 0 to a


be a
movable point P and a a given constant vector and consider the linear
vector equation
r •a = c , (14-3)

in which c is a scalar constant and the quantity on the left side is the scalar
or direct -product of the vectors r and a.1
Explicitly,
r •a = xax + yay + za, = c = const . (14-4)

Equation (14-3) or (14-4) is the equation of a plane, and since the projec
tion of the variable space vector r = OP on the con
stant vector a times its absolute value must be equal
to a constant c, this plane must be perpendicular to the
vector a.
Again, if vector r = OP is an independent variable,
consider either of the two vectors v and r' as dependent
variables. Draw the new position vector r' = UP' from
the origin 0 to a second, movable point P' and denote
the vector originating in point P and ending in P' by
v =TP' (Fig. 14-2). Then
v = r' -r . (14-5)
FlO. 14-2.

Suppose that v is defined by the vector equation

v = (r •
a)b (14-6)

in which the quantity appearing in the bracket is the scalar product of r


and a. a and b are given constant vectors. The parentheses may as
well be omitted and this vector equation written more simply as follows:

v = r • ab (14-7)

r • a quantity, which has only magnitude but no direction,


as a scalar
represents a factor with which vector b is to be multiplied. Thus vector
v must be oriented in the direction of and be parallel to vector b. From
Eq. (14-3) we know that, if r • a = const, point P must belong to a plane
The scalar product of the vectors a = iax + jav + ka, and b — i6* + j6» + k&«
1

is written a • b. If ^ is the angle smaller than * between the directions of a and b


<p

the scalar product is denned by a ab cos <por cos or equal to the projection

i •
|a|

1, |b|

of a on times the absolute value b. Since • = • = etc., using the distribu


0,
b

j
i
i

tive law when carrying out the multiplications we obtain


■ = axbz + + aj>, . (14-3o)
b
a

<zv&„
Also
• = • .
b

b
a
a
154 THEORY OF FLOW AND FRACTURE OF SOLIDS

E (Fig. 14-3) to the fixed direction of vector a.


which is perpendicular
We therefore see that vector v for all points P situated in such a plane E
has a constant value and is oriented in a direction parallel to vector b.
We may interpret vector v = PP' as the displacement of point P in the
strained condition of a substance in which a material point P moved in a
new position, P'.1
If, for example, the two constant
£—1—— vectors a and b are parallel, v also has
the same direction and Eq. (14-7) rep
/ 1 resents a uniform extension of a material

-V?
/ in the direction a. [r a is constant

in all planes which are perpendicular


to a and increases proportionally to the
product r cos (r,a) or with the length
0*
Fio. 14-3. Displacement vector v. of the normal to these planes drawn
from the fixed point O.]
Similarly, if b is perpendicular to a we obtain a displacement vector v,
which in the planes normal to a has the same value which is proportional
to the length of the normal to them drawn from the fixed point O and
which points always in the same direction perpendicularly to a. Equa
tion (14-17) evidently represents a simple homogeneous shear.
Thus we see that a linear vector function such as Eq. (14-7) may serve
to represent states of finite homogeneous strain if the vectors v are
interpreted as the displacements of the points of a substance.
A displacement vector v can also be defined by a vector product2
v = a X r (14-8)

where r and v are the independent and the dependent variable and a is a
constant vector.Draw vector v from the end point P of the variable
position vector r = OP. According to the definition of a skew product,
1 It is easy to see that this is in general a certain linear homogeneous distortion and
that parallel planes are displaced in parallel planes, parallel lines into parallel lines,
etc. This was described for a homogeneous state of finite strain in Chap. 11.
* The vector or skew product of a into b is the vector which is
perpendicular to the
plane of both, has the absolute value ab sin <p,and points in that direction in which the
third finger of the right hand points if the thumb and forefinger are coinciding with
the directions of a and b in this order. Since i X i = 0, i X j = k, etc., the skew
product is defined by

J i(a„b. — aj>,)
a. \
a X b = a. = +}(a.b, aj>.) \ (14-8a)
+k(a,b a„6,) . )

We note that
a X b b X a.
VECTOR GEOMETRY OF STRESS AND STRAIN 155

v must be perpendicular to the plane of a and r and its absolute value


= = ar sin <p must be proportional to the distance of point from

P
|v|

.
the straight line in which lies. The linear vector function expressed

a
by Eq. (14-8) represents the distribution of the velocities v in a rigid
body turning around a fixed axis coinciding with the position of the vector
a.
A linear vector function
expressing displacement v may contain two

a
or more additive terms similar to that introduced in Eq. (14-7)

:
= •
aibi •
+ • • • .

+
a2b2 (14-9)
v

r
The At, an . . . bi, b2, . . . are constant vectors. The distortion rep
resented by the first term alone in Eq. (14-9) leaves the plane passing
through the origin perpendicular to the vector at in an undistorted
O

condition. If the right side of Eq. (14-9) the sum of two terms like

is
those which have been printed explicitly in Eq. (14-9) this vector func
tion in general leaves the straight line in an undistorted condition in
which the two planes passing through which are perpendicular to the
O

vectors at and a2, respectively, intersect each other. If alf a2, and a3
in Eq. (14-9) are three noncoplanar vectors we see that the most general
linear vector function representing distortion of body in which no
a

a
straight line remains unstretched must be expressed by three such terms.
This sum of three terms:
= •
(&ibi + asb, + a3b3) (14-10)
v

may be abbreviated by
= r*>
v

* .J (14-11)
1

where = aibi +
+

a2b2 a3b3

pair of vectors such aibi according to nomenclature introduced


A

as
by Gibbs called a dyad; sum of dyads such as a dyadic (binominal) *
is

(or an "affinor"); ai,a2,a3 are the antecedents, bi,b2,b3 the consequents.


Pairs of vectors ab without the product sign of dot or a cross between
a

them have a meaning preceded or followed by dot or cross con


if

necting the dyad with a third vector (such as r). The order in which
r

and 4> appear essential. In the dyadic *


called postfactor, *
is
is

a
r

in *
• prefactor. aibi and biai are called conjugate dyads,
a
r

* = aibi
+

a2b2 a3b3

and *c = b^i + b2a2 +


are conjugate dyadics.
b3a3 The symbol 4> in
fact denotes an operation, not quantity in itself. Since
a


a)b = r)b
(a

(r
156 THEORY OF FLOW AND FRACTURE OF SOLIDS

and this is the same as b(a • r), it follows that


r • * = *c • r . (14-12)

To visualize the meaning we can attribute to such an operator * we

quote two examples. Assume

v = r • * (14-13)

where the dyadic is given by its expression in Eq. (14-11).


4» Carrying
out the mathematical operations symbolically indicated on the right
side of Eq. (14-13) amounts geometrically to comparing the original
positions of points P in space with their new positions P' (actually con
structed by adding geometrically the vectors v = PP' to the vectors
r = 0~P describing the original positions of P). It has already been
stated that by such a process we can reproduce the displacements in a
material under homogeneous states of finite strains, such as pure exten
sion or a simple shear. Thus the linear vector function defined by
Eq. (14-13) can serve to represent homogeneous states of strain.
Suppose that r = UP is a variable position vector in space, W a given
diadic and that a second vector r' = OP' is drawn from the same fixed
point 0 defined by
r' = r • V. (14-14)

Such a linear vector function may serve to express what is termed a linear
transformation in space. We may state that multiplication of vector r
by operator *8* turns the variable vectors r into new positions r'. This
interpretation of Eq. (14-14) is suitable also for expressing the rotation
of rigid bodies.
The necessary and sufficient condition that Eq. (14-14) should express
a rotation of a body without a distortion is that the diadic V is reducible
to the form
*F = ii' + jj' + kk' (14-15)

in which the antecedents and the consequents represent sets of mutually


perpendicular unit vectors. If this operator W is applied to vector
r = ix + iy + kz the new vector r' = i'x + )'y + k'z results, which has
the same components as r had along the new directions i',j',k\
A special case of Eq. (14-15) is the diadic
ii + jj + kk = I (14-161

which is called an idem factor and denoted by I. The scalar product of a


vector r and an idem factor is equal to the vector r
r.I = I.r = r. (14-17)
VECTOR GEOMETRY OF STRESS AND STRAIN 157

Thus I as a pref actor or a postf


actor leaves the vector r unchanged.
The two geometric interpretations of the scalar product of a variable
space vector r with a dyadic [Eqs. (14-13) and (14-14)] may in fact
coincide with each other. The linear space transformation [Eq. (14-14)]
and the homogeneous distortion [Eq. (14-13)] describe identical operations
if the displacement vector in Eq. (14-13) is made equal to
v = r' -r , (14-18)

where r' denotes the vector which was introduced in Eq. (14-14).
After
substituting for v and r' their values given in Eqs. (14-13) and (14-14)
and after making use of the identity r = r • I, from Eq. (14-18),
r - * = r •
(V - I) ,

it is seen that the dyadic *V is

»F = * + I. (14-19)

Consider a linear transformation


r' = r • T. (14-20)

Let the antecedents ai,a2,a3 and consequents bi,bi,ba in the dyadic


VP = aib! + a2b2 + a3b3 (14-21)

represent in each group a system of three noncoplanar vectors originating


from a fixed point O (Fig. 14-4). The group ai,a.,a3 defines a reference

Fio. 14-4. Flo. 14-5.


Fios. 14-4 and 14-5. Affine transformation.

system in space. Geometrically it may be represented through the three


straight edges of a corner of a solid parallelepiped inclined under given
angles. Suppose we choose r = at, in Eq. (14-20). After scalar multi
plication a new vector Ti is obtained:
r/= 01^1+ aiO2 cos ^ub2-|- aid3 cos ^nbs. (14-22)
158 THEORY OF FLOW AND FRACTURE OF SOLIDS

The edge ai has been changed in the new vector T\. Similarly we
show that the edges 82 and a3 are transformed in the new vectors r2' and
r3'. Thus we see that the linear vector function [Eq. (14-20)] causes the
edges ai,a2,a3 of an oblique parallelepiped to transform in the new edges
li'fijt* of another parallelepiped.1
For the antecedents we may conveniently choose unit vectors without
loss in generality. If ai,a2,a3 denote three arbitrary unit vectors in space,
we obtain from Eq. (14-22) more simply:

r/ = bi + cos v>u b2 + cos ipn b3 1

r2' = cos <pu bi + b2 + cos <p2s b3 > (14-23)


r3' = cos tp3i bi + cos ^>32b2 + b2 J
If these unit vectors coincide with the mutually perpendicular unit
vectors i,j,k of the x,y,z axes of a cartesian system, we obtain
x/ = r • V -r •
(ibL + jb2 + kb3) , (14-24)

which transforms a unit eube (Fig. 14-5) in an oblique parallelepiped:

Ti' = b! , r2' = b2 , r3' = b3 . (14-25)

It appears that it is thus convenient to normalize a dyadic by choosing


i,j,k for its antecedents.
14-2. Nonion and Normal Form of a Dyadic. Let
* = ibj + jb2 + kb3 (14-26)
be a dyadic where i,j,k are perpendicular unit vectors and
bi,b2,b3 three
noncoplanar After substituting here the components of each
vectors.
vector, for bi = ifru + jfei2 + k6i3, etc., the dyadic m^r be written in
its nonion form. The nine dyads of which it is composed are listed
in order as follows:

buii + buij + 6nik


+62iji + 622jj + Mk .
(14-27)
l+63iki + 632kj + 633kk
This is frequently abbreviated by omitting the dyads and
ij,
ii,

. . .

by listing only the constants in their proper location in matrix:


a

bu bn bn
4> = •
621 622 623 (14-28)
b3i 632 633

Remembering that linear transformation parallel lines or planes again


1

changes
a

into parallel lines or planes.


VECTOR GEOMETRY OF STRESS AND STRAIN 159

The nonion form of the conjugate dyadic *c = bii + + bak is


D2J

bti 631I

bit

(bn
2>22 63s! (14-29)
b\3 bi3 633J

in which the rows now contain the components which were formerly
assembled in columns, and vice versa.
A dyadic symmetric does not change its value upon replacing

if
is

it
the rows by the columns, i.e., *c = antisymmetric *;

is
hereby

if

it
it

if
changes its sign, i.e., *c = — *• Any dyadic can be split in the sum *
if

of symmetric part H(* + *c) and an antisymmetric part 3-i(* — *c)


a

:
- ^{bti +26n 6u

+
621

H(*
f

+ *e) fcu 2bM


631 + bn ba + bn
(14-30)
bu — b2\
- *e)
0

m+ =
bu — 612
0

,"31
— —
O23
013 632

Thus any linear transformation:

u = 6nx
+ b21y + b3lz
,

= 612X
+ bay + bzig
v

w = bux + bi3y + b33z

corresponding to the linear vector function v = * represented in


is

r

one and only one way by the sum of o symmetric part:

, +
+

(612 621) (&13 b3l)


bux s
H,

H,
y

o
z
>

(621 + 612)
x + bity +
(b" + M
z

(631 + b\3) (632 + 623)


,

x +
+

.
1

!3Z
2

and antisymmetric part:

(621

612) (631

613) ,
,

,
0


(6.2 -62.) (632 623)
x + +
.

o
z
2

(6.1 -631) x+(bu-btt)y + Qt


160 THEORY OF FLOW AND FRACTURE OF SOLIDS

corresponding to
~
<* +
and r .
(» »c) ,
2 2

respectively.
Examples of Dyadics: Let the vector function v = r • ♦ represent dis
placement vectors in space. If *
= «ii, we obtain a pure extension in the

direction parallel to the x axis for conventional strain «(— 1 < e < oo).
We obtain three pure extensions in three perpendicular directions if

* = eiii + ejjj + ejrk ,

a simple finite shear in the direction parallel to the y axis if

<I> = -yij •

If c is an infinitesimal quantity, a small rotation around the z axis is


represented by
* = c(ij - ji) .

If c is an infinitesimal quantity, a small pure shear without a rotation is


represented by
$ = c(ij + ji) .

If ci,c2,c3 are infinitesimal quantities, the dyadic for an infinitesimal


rotation of a rigid body around a certain axis is represented by

0 ci c2
— 0
ci c3

-ci — c, 0 )

We have seen that a dyadic *


causes the distortion of a cube in an
oblique parallelepiped. Inscribe a sphere of unit radius in the unit cube
i,j,k assumed in Sec. 14-1 [Eq. (14-26)] using the origin O of the coordi
nates x,y,z for the center of the sphere. The transformation defined by
Eq. (14-26) causes the distortion of this sphere in an ellipsoid. Three
radii in mutually perpendicular directions in the sphere will become a set
of conjugate radii in the ellipsoid. Since the principal axes in the ellips
oid are such a set, we see that three originally mutually perpendicular
directions must exist in the unit sphere which after the transformation
will become the principal axes of the ellipsoid. We conclude that a
dyadic
* = ia + jb + kc (14-31)

can be expressed in an equivalent "normal form"

* = ciaibi + c2a2b2 + c3a3b3 (14-32)


VECTOR GEOMETRY OF STRESS AND STRAIN 161

in which the antecedents and consequents are unit vectors in mutually


perpendicular directions and Ci,c,,cj are scalar constants. A special
example has been quoted in = *
tiioio + c2jojo + «akoko representing a
finite homogeneous strain in which the principal directions io,jo>ko do not
rotate while the dyadic *
defined in Eq. (14-32) assumes a rotation of the
principal axes. Distortions of this
kind are known also as "affine" trans o(a„ay,az)
formations (of a cube or of a sphere).
The ellipsoid just mentioned is the
strain ellipsoid if Ci,Cj,cs express unit
strains.
14-3. Stress Tensor. Consider a
general state of stress. Let ax, <r„,
<r*, t„,, t,x, Txy be the components of
stress in the sections parallel to the
Fio. 14-6. Stress tensor.
coordinate planes and denote by a
and t the normal and shearing stress in an oblique plane section normal
to a unit vector a :
a = ia* + ja„ + ka« (14-33)

so that the components of a(a1,a„,a,) are the direction cosines of the


normal to this plane section. They may be interpreted (as in Chap. 10)
as the coordinates of the point of tangency of a plane tangent to a unit
sphere. The components sx,sy,s, of the resultant stress s
s = isx + js„ + ks, (14-34)
in the oblique plane are [Eq. (9-1)] equal to

8X
= OVZj + Tyxav + 7,1a, ,

Sy = Tiyax + Cydy + TZydt , (14-35)


8, = TrA + Ty,dy + Ord, .

The coefficients in these three linear equations for az,av,a,


<rx, ay, .

satisfy the symmetry conditions t^ = tvx, t», = t2V, tzx = txi. The stress
components ox,rm,Tm, are the rectangular components of a vector S*
(a resultant stress). S* and similarly S„,S, (Fig. 14-6) are defined by
the equations

Sx = lO-x + JTxy + kTx


= +
S„ ir„x + jo-v kT„ (14-36)
S* = iT„ + + k(T2
JTx»

The three Eqs. (14-35) in fact define a variable stress vector s in space as a


162 THEORY OF FLOW AND FRACTURE OF SOLIDS

linear function of a second variable vector a and can be condensed in the


linear vector function:
s = a • n (14-37)

with the dyadic

n = iS, + jS„ + kS, . (14-38)

a is the independent, s the dependent variable vector; the six vectors


(ijjkjSoSyjS,) constituting the dyadic n are given constant vectors.
After substituting the values of S^S^S, from Eqs. (14-36) in Eq. (14-38)
the nonion form for n :

<T, Tzy T,2

II = Ty, <ty Tyi (14-39)

is obtained.
We note once more that a stale of stress [Eqs. (14-37) and (14-38)] is
defined by a symmetric dyadic n in which six components can be freely
chosen and in which r„ = tv,, tv, = rn, t„ = t„. Because a state of
stress can always be reduced to its three principal stresses, i.e., to three
pure tensions (or compressions) <r\,ai,<r2 acting in three mutually perpen
dicular (the principal) directions (see Chap. 9), symmetric dyadics such as
n have been called "tensors."1 [Gibbs and Wilson call them "right
tensors" ("Vector Analysis," p. 351).]
The normal stress a acting across the oblique plane section normal to a
is the projection of the resultant stress s on the normal a or

<r = s. a = a. n. a. (14-40)

After substituting here the value of the stress tensor n from Eq.
(14-38), using Eq. (14-36) and after carrying out the scalar multi
plication, we obtain for <r a homogeneous second-degree expression
of the direction cosines a,,Oy,a, which was derived in Chap. 9 Eq.
(9-4):

a = a,a,2 + <ryi„2 + <r,a,2 + 2T,va,Oy + 2rv,a^a, + 2Tt,ata, . (14-41)

The six components of the tensor n appear again in this homogeneous


second-degree function of a,,ay,a, as coefficients. These six coefficients
<T,, vy, <Tt, Tiy, Tyt, t,, are called "tensor components."

'The term "tensor" as just defined was suggested by the physicist Woldemar
Voigt in Gottingen in his work on crystal physics. The term tensor has since been
interpreted in many different senses.
VECTOR GEOMETRY OF STRESS AND STRAIN 163

If the coordinate axes are made to coincide with the principal direc
tions, the stress tensor n is expressed by

n = iSi + jS2 + kSs = o-iii + 0-2JJ + o-3kk . (14-42)

It has the matrix

n =
(14-43)

where |Si| = o\, |Ss| = a, |S3| = <rs are the principal stresses. The expres
sion for the normal stress a in this case reduces to

a = a\(iz2 + <rtfiu- + <T3<z,2. (14-44)

It was shown in Chap. 9 that the principal stresses can be computed


from Eq. (14-41) by making the normal stress <r an analytic extremum.
The coefficients of the resulting cubic equation were invariants of the
state of stress. The first invariant was equal to the sum of the three
normal stresses <rx + ay + at. Thus the mean stress

+ <Ty + <T,
9 = (14-45)

is one of the invariants of the stress tensor n. After subtracting the


value of the mean stress from the normal stresses of tensor n the part
9
of the state of stress is obtained which was called a deviator. The
deviator of tensor n is

— 3

D = ffy — 9 (14-46)
— 9
<r* )

14-4. Finite Strain. Referring to Sec. 14-1, Eq. (14-24) suppose that
i,j,k are unit vectors coinciding with the directions of rectangular axes
x,y,z and i'j'jk', a triple of unit vectors in given directions. Define

i' = aizi + aty) + axtk ,


j' = aVI\ + dyy] + a»Jt , (14-47)
k' = a„i + azyj + a„k .

The components of the matrix:

(14-48)
".-.-
164 THEORY OF FLOW AND FRACTURE OF SOLIDS

are the direction cosines of these three straight lines coinciding with the
unit vectors i',j',k' and subject to the conditions

+ a,y2 + a,,2 = 1 , ay,2 + + Oy^ 1,


a,,2 =
(14-49)
o«2 + a*i/2 + 1 .

Assuming that the vector r = OP = ix + jy + kz defines the original


position of a point P in the unstrained condition of a body and that the
vector r' = OP' = ix' + jy' + kz' defines its position in the strained
condition of the material, as we saw, a state of homogeneous finite strain
can be expressed by the linear vector function

(14-50)

if the dyadic is chosen so that Eq. (14-50) transforms the originally


mutually perpendicular directions i,j,k into the three given oblique direc
tions i',j',k' of Eqs. (14-47), respectively, while simultaneously the body
is stretched along the latter directions by given amounts. Let \,,\y,\,
be the prescribed elongations in the directions i',j',k', respectively. An
elongation according to Chap. 12 is defined by X = (1 + «)2 where e
denotes a conventional strain. The dyadic causing this strain according
to Eq. (14-24) is

* = Vx^ii' + V^vH' + Vx^kk' (14-51)


having the matrix
\/\, a, \/x^ a*
y/\y ay \/\y ay , (14-52)
a.

After substituting Eq. (14-51) in Eq. (14-50),

r' = y/\xxi' + VK y)' + V£ zk' (14-53)

is obtained, which, after the components are separated along the x,y,z
axes using Eqs. (14-47), furnishes the equivalent system of the linear
transformation expressed in rectangular coordinates:

x' = \/\, a,,x + y/\v Oy,y + \AI a,xz ,

y' = \/K a,llx + amy + y/\, a,yz , (14-54)


z' — y/\, a,,x + y/\y Oy,y + a„z ,

with its constants fitted to the conditions which were imposed on the
strain, namely, that a cube having originally the edges ri,rj,rk is distorted
in an oblique-sided parallelepiped having the edges

a = ri' , b = y/\ rj' , and c = -\/\, rk' ,


VECTOR GEOMETRY OF STRESS AND STRAIN 165

respectively. Dyadic * [Eq. (14-51)] is also expressed by


* = - (ia + jb + kc) . (14-55)

This strain causes an elongation X in an arbitrary direction, which


shall be computed now. It is found from the equation

r*X = r" = r' • r' = (r •


*) • (*e •
r) . (14-56)

The scalar product of the dyadic * with its conjugate <Dc

*c = - + bj + ck)
r (ai (14-57)

is again a dyadic. Let it be denoted by V. Obviously V has the non-


ion form

fa
• a a • b a - c

.= -£ b-a b.b b.c •


(14-58)
r • a c • b c • c
lc
Using from Eq. (14-56) we have for

r" = r • iy • r (14-59)

and note that while *


in general had nine unequal components the dyadic
according to Eq. (14-58) is symmetric and has only six unequal com
ponents. Thus the elongation X = r"/r2 in a given direction depends
on the tensor *F = • *
^c.1

In view of many useful practical applications which may be made of Eqs. (14-58)
and (14-59) for the analysis of finite strains it is worth while to evaluate the scalar
products appearing in the six components of tensor V in Eq. (14-58) explicitly.*
We find
V'XiX, cos a,, v'X,X, cos ax,\
V X„X, cos
IXS \/X|fX* cos ctyi X„ a,. > (14-60)
\/X,X, COS a„ "y/X/Xy cos a„ X, I
1 The perfect analogy of the vector forms
a -r • H •r and r" = Xr2 -r - «F • r

expressing the normal stress a in a state of stress and the square of the radius vector r' in
a stale of finite homogeneous strain is noted, a and r" = Xr* are both derived from
pure tensors (n and V, respectively).
* Since a =
-\/xI tV and b = \/x7 rj' their scalar product is
a •b = VXiX„ r2i' • j' = v/x.X,, r* cos «,„

where a,y denotes the angle under which the unit vectors i' and j' are inclined with
respect to each other. Similarly the products b • c, c • a furnish the terms in the
matrix [Eq. (14-60)].
166 THEORY OF FLOW AND FRACTURE OF SOLIDS

and after evaluating the right side of Eq. (14-59), inserting in it these components of
tensor V, and substituting for the ratios x/r = a,, y/r = ay, z/r — a, the direction
cosines a,,av,a, of vector r = 01', the elongation X = r''/r* equal to

X = XiO,1 + XjO,1 + X«a,1 + 2 VXi,X« cos ay/iyd, + 2 Vx.X, cos n_,a,fiz

+ 2 \/XxX„ cos o,,o/is . (14-61)

The six coefficients in this second-degree expression of the a„av,a, for the elongation
X in an arbitrary direction under a finite homogeneous strain are just the six com
ponents of tensor W.
Love" has developed what amounts to an equivalent expression of Eq. (14-61) of a
quantity equivalent to elongation X. We shall express the quantities introduced by
Love in his analysis of a finite homogeneous strain by means of the components of
tensor V. Love starts by considering a linear transformation in the form

x' = (1 + c„)x + c„y + c,jl , >


y" = cy,x + (1 + cv,)y + c,»8 , > (14-62)
z' = ex
+ c,vy + (1 + c„)z . )

Let us compute the value of the elongation X from the ratio

r" x* + y" + z"

by using the preceding equations and the direction cosines a, = x/r, ay = y/r>
a, «■ z/r. We find the expression of second degree in the cij,a,,a(:

X - - 2[c«
1 + HicS + e„* + c,*J)K« + • • •
• ■■.
+ 2[c», + C» + cyrCyV + c,yCxx + CxyCI,\aya, + (14-63)

The six constant terms contained within the brackets (of which only two were
explicitly written out) are the ones which Love denoted by the symbols: «»,, <„,, ««.,
«»•, ««*, «i» where
«*« = c„ + H(c«» + c,,» + c„«) , ■• •
(14-64)
tyc — Cyz I Cgy T CyffCyy ~\~CgyCzx 1 CzyCzZ , * * * •

After substituting these symbols in Eq. (14-63), he obtained

X — 1 = 2(«„aI1 + tyytty'1 + «„a,* + f„.-a.,o- + t.zdji, + tzra,a,) (14-65)

Comparison of this with Eq. (14-61) shows that Love's six quantities are equal to

fix = H(X» — 1) , €v, =VX^XT cos orVI , \


tyy = }-z(K — 1) , e„ = VX,X* cos a,* , > (14-66)
«»« = H(^»

1) ) <:>
= VX*X» cos aXy . I

Neither the €„, «,,, «,, nor the «„,, e„, «,v express in a strict sense "components of
strain" (except for an infinitesimal strain).
14-5. Nonhomogeneous Finite Strain. Now that we have described the simpler
kind of vector fields in which the three components of a field vector were assumed as
linear functions of the three components of a second vector, let us consider the general
case when they are functions of the latter. Familiar examples of vector fields of this
1 Treatise on the Mathematical Theory of Elasticity," 4th ed., p. 66, Cambridge
University Press, London, 1927.
VECTOR GEOMETRY OF STRESS AND STRAIN 167

nature are the displacements of the points of deforming bodies, the velocities in a fluid
in motion at a given time, etc. Before treating the case of finite nonhomogeneous
strain means must be developed for the differentiation in a vector field.

jt>
Suppose that a vector v = iu + + ku',1 the three components u,v,w of which are
functions of the components x,y,z of a space vector = ix +

jy
+ kz, increases by

r
the differential
dv = — dx +
T — dvy^
+ — dz (14-67)
ax dy dz

when changed by an increment dt = dx + dy + dz. After writing hero for


is

k
r

j
i
dx = dr • dy = dt • dz = dr • and after rearranging terms we obtain

k
j,
i,

v+dv
=
dl-Vax+>Ty+k6z) (14-68)

Obviously the three additive terms within the parentheses define


a dyadic. Hamilton introduced the symbol for

it
. dv . dv dv
,

,
,
dx dy dz
du . dv dw\
idx + ka7) +
+,

,
,

,
(14-69)
(,

iax
remarking that the differential operator V, called "nabla'
by him,

_ .9 . . . ,
J

(14-70)

behaves like a vector in vector equations, Thus the increment dv of vector v in


abbreviated writing
dv — dr • Vv (14-71)

(pronounced dr dot nabla in fact a linear vector function, like


is
v)

=
jb

• 4> =
(ia + kc)

+
v

(14-72)
r

which was treated previously. Since dv/dx = iu, + jv, kw„ (the sub-
+

scripts x,y,z when attached to the components u,v,w of vector v designate here partial
differentiations with respect to x,y,z) the nonion forms of Vv and of its conjugate
defined by
dv dv . dv .
vV
,

,
.

dxl+dy>+Tzk
are given by the matrices:

v, w,\ iv, u„
U, B, It, VV = IV, V, (14-73)
{u,

Vu
,

.
J

u, W,l w,i
'

V, W, Wy

The symmetric (tensor) part of dyadic Vv J^(Vv vV), the antisymmetric part
+
is

H(vv —
w).
The symbol one of the three rectangular components u,v,w of vector v, should
1

v,

not be confused with the absolute value of vector v, namely, = Vu* + »* w*.
|v|

+
1G8 THEORY OF FLOW AND FRACTURE OF SOLIDS

Divide Eq. (14-71) by the absolute value dr of the vector increment dr. di/dr is a
unit vector in the direction in which the increment dr was taken. Denote it by a.
Then also

dv
a • vv . (14-74)
dr

The space derivative of a vector v in the direction of a unit vector a is equal to the
scalar product of unit vector a and dyadic Vv.
The space derivative of a scalar quantity (such as a temperature, for example) is
expressed in similar manner. Let U be a scalar function of the coordinates x,y,z.
Then

dU _ all dx dU dy dU dz
(14-75)
dr dx dr dy dr dz «r

This equation shows that dll/dr is the projection of a vector (the gradient of U) on
the direction in which the increment dr is taken. The gradient in a scalar field is a
vector, which is oriented in that direction normal to the surface U = const in which
U increases and has the absolute value

<- '" -
V(£)'+(£ )'+(§)■ (14-76)

Let a = di/dr be a unit vector in the direction of dr. Then

all .all
hi' —dy— ,H
aU = a • (. k aU\
, . — , TT
• grad U
J
—— —
) = a

( i
dr dx \ dz
6 (14-77)*

* While the derivative dv/dr of a vector v depends on the dyadic Vv, dU /dr in a
scalar field is the component of vector grad U in the direction of a = rfr/dr.
Two rules are quoted for the nabla operator:

a , , a , . a\ ,, . . . , . au , dv , dw ,.
+ +>+■»>
(.ldx+idy kTz)-
9« -d-x+dy+Tz'dlVV
i i k
d a a
V X v
(ll+la\+*al)*^+t° +
k")-£dx dy dz
a V IC

dv\ , . (du dw\ . . ( dv du\


Ty-Jz)+1\Tz -to)+kU"^)=r0tV-
(dw
(14-78)

The first is a scalar quantity which is called the divergence, abbreviated by

V •v - div v
du . dv
+
dw
(14-79)
dx dy dz

while the second defines a new vector called the rotation of the vector v, abbreviated
by rot v (equal to twice what is called the curl v). If vector dA denotes an infinitesi-
mally small area in space and v the velocity of a fluid streaming through this area

v • dA = v • (i dAt + j d.4„ + k dA,) - u dAx + v dA, + w dA. (14-80)


VECTOR GEOMETRY OF STRESS AND STRAIN 169

Returning to the treatment of states of nonhomogeneous finite strain assuming that


the components of displacement are functions of the coordinates x,y,z, let u = x' — x,
v = y" — y, w = z' — z be the components of the displacement vector v. If r is
changed by an increment dr, vector v according to Eq. (14-71) changes by the increment

dv = dr - Vv (14-83)
where
ux Uy u,
Vv =
\ v, v,
Vy V,
v,
\ (14-84)
W, Wy Wz

the subscript* of u,v,w denoting partial differentiation.


A finite homogeneous strain was expressed by the linear transformation
v = r,—r=r.P (14-85)
where
?xx C,y C,i I
• „. -I-„ Cyy Cy, \ . (14-86)
,c. c,, <•„)

Equations (14-85) and (14-86) are equivalent to the three linear equations (14-62),
page 166, from which the quantities «x,, «,„ . . . introduced by Love were computed
in Eq. (14-64). By comparison and analogy we obtain the expressions for the quanti
ties t,,, tv,, . . . which were introduced by Love for a nonhomogeneous finite strain
by substituting in the formulas Eq. (14-64) instead of the components c„, cai, of ...
dyadic * the components u,, ».,...of dyadic Vv:

t,, = U, + }4(u,2 + V,' + W,2), - - -


= «'„ + V, + Uyu, + VyV, + WyW„ . . .
«„f (14-87)

is the amount of fluid pushed through the area dA. Let dA, = dy dz, dAv = dz dx,
dA, = dx dy and consider the small volume element dx dy dz. Then the double
integral of v - dA taken over the entire surface of the element dx dy dz is

J JvdA =
(~;+^ + ^dxdydz
= div vdxdydz. (14-81)

If this last equation is integrated over the elements dx dy dz within a space bounded
by a closed surface, the contributions to the left side of Eq. (14-81) will cancel out for
all internally situated small surface elements dA. Only the elements dA of the bound
ing surface will add contributions. Equation (14-81) applied to a finite volume
bounded by a closed surface thus gives

div v dx dy dz = jJ v dA (14-82)

expressing the theorem of Gauss. The integral on the left side is to be taken over
the volume, that on the right side over the closed surface.
If u,v, w are infinitesimal quantities and the components of a displacement vector
v the quantity div v expresses the dilatation of the volume per unit volume. Since
v • dA measures the volume which is pushed through a surface element in the direction
parallel to v, div v for a finite strain does not express the true dilatation of volume per
unit volume.
170 THEORY OF FLOW AND FRACTURE OF SOLIDS

If ds denotes the element of arc of a curve drawn through point P(x,y,z) in the
unstrained condition and ds' the length of this element in the strained condition of
the body, the elongation X in point P'(x',y',z') in the direction of ds' in a nonhomo-
geneous field of finite strain is defined by

-m (14-88)

and X is found after substituting in Eq. (14-65) the expressions of Eq. (14-87), for
i,,, «,,, ...
in terms of the partial derivatives uz, . . . and for the direction cosines
a, = dx/ds, a„ = dy/ds, a, = dz/ds.
14-6. The Condition of Incompressibility of Material. The Affine Transformation
of a Sphere of Such Material. Let a state of homogeneous finite strain be given by
Eqs. (14-50), (14-51), and (14-54). Suppose that we have to deal with an incom
pressible material. A tetrahedron of material the corners of which in the unstrained
condition have the coordinates x,y,z = (0,0,0,), (1,0,0,), (0,1,0) and (0,0,1) and in the
strained condition x',y',z' = (0,0,0), {xa!,Va ',zA'), (xB',yB',zB') and (xc',yc',zc')t
respectively, has originally a volume equal to V = }4 and in the strained condition:

1 0 0 0
Xa' VA ZA
xa'
V =
1
1
_ 9
■TH
.. /
Va ZA
_ /
= H Xb
.
Vm zr (14-89)
</H ZB
, t Xc yc tc'
1 Xc !lc' zc

The unit vectors r = i, j, k become the vectors r* = V^ V^i V%k' after the

l'i

i'i
strain applied. The volume V therefore also expressed by
is

is

V (14-90)

For incompressible material V =


V,

therefore:

V^iKK X) = (14-91)
1

where 3) denotes the determinant in Eq. (14-90)


For an incompressible material the affine transformation a sphere on account of the
of

condition takes a simpler form. sphere


A

(14-91)

z» = rs
y*

+
+

Xs (14-92)

embedded in such a material becomes an ellipsoid of equal volume having the con
= r(\/\t i',
\/\„ \/yi k'). The equation of this ellipsoid obtained
j',

jugate radii
r'

is

by solving the three linear equations (14-54) given in Sec. 14-4 for x,y,z:

T' a,* a„
=
Z7 ;/'

avv aty a„
y
,

a„, a„| On
a,. x1
(14-93)
= Vk a„
K

V1
z

a« n ».■
z

by squaring these last three expressions, adding them, and making their sum equal
to r».
VECTOR GEOMETRY OF STRESS AND STRAIN 171

The resulting equation when r is taken equal to 1 is the equation of the strain ellipsoid
for an incompressible material.
14-7. Correspondence of Direction Cosines in Unstrained and Strained Condition
of Material. We shall only indicate the means for establishing the correspondence
formulas expressing the direction cosines a, = x/r, ay = y/r, a, = z/r by a,' = x' /r',
ay' = y1 /t' , a,' = z'/r' or vice versa.
Divide Eqs. (14-54) by r' on their left and r' = \/\ r on their right sides. This
gives

°I' =
(Vxl a„a, + \/\^ a„,a, + Vxi a„o.) ,

°,' = (y/K a,ya, + Vxi o„„a„ + Vx^ a,va.) , (14-94)

°x' =
~7=
(VxTa.ra, + Vxjl a„,a, + \/x7a..a.) .

After squaring these, adding and equating the sum equal to 1 we find a complete
expression for the elongation X = r'*/rx [equivalent to Eq. (14-61)]. Substituting X
back in Eqs. (14-94) the formulas expressing the a,',ay'a,' in function of the a,,ay,a,
are obtained.
The corresponding procedure carried out on the set of the three linear Eqs. (14-93)
furnishes the reciprocal formulas expressing the a,,ay,a, as a function of the a,',ay',a,'.
Divide Eqs. (14-93) by r on their left and by r = (l/\/x)r' on their right sides. This
gives
|O,' a,, a,:
a, — Vx Vx„x, Qfyy On (14-95)
[a,1 a„, a„

After squaring, adding, and equating the sum equal to 1 we obtain the expression
for the reciprocal of the elongation or for 1/X as a function of the a,',ay'a,' and after
substituting this back in Eqs. (14-95) the formulas for the a,,ay,a,.
14-8. The rules laid down for the denotation of the components of the tensors of
stress <r,, <ry, <r„ Tye, Ti,, T,y and of strain ex) e„, t„ yy„ y,B, y,y in cartesian coordinates
as developed in the preceding five chapters seemed to the author to be in general
usage in the best books on engineering mechanics of continuously distributed masses.
The superb symbolism of the dyadic calculus of Gibbs relates their geometry to the
theory of linear vector functions.
No use is made in this chapter and in other parts of this book of a symbolism of
abbreviations of different character which has made its appearance recently in the
engineering literature under the name of tensor notation and which disregards well-
established ways of writing out all the terms in tensor equations or in relations con
necting components of stress and strain, etc. The author questions the value of this
new symbolism in the theories of elasticity and plasticity and from the engineer's
point of view for a number of reasons. Space prohibits enumerating them here, but
the primary reason is that it does not introduce any new concepts apart from its
apparent brevity, alleged by a few of its adherents, which evanesces upon an objective
examination when the new notation is compared with the one in common use for
components of tensors and in view of the fact that the denotation of classic origin
dating back to Cauchy and accepted by the best authorities in mechanics has been
considered adequate by all these workers during more than a century for distinguish
ing the components of stress or strain relative to their directions.
172 THEORY OF FLOW AND FRACTURE OF SOLIDS

Apart from the novelty, no new facta have been uncovered to demonstrate that
the established way of expressing the symbols of stress or strain is inadequate or
impractical, the truth being just the contrary, because in most applications of the
theory of elasticity or plasticity in solvable cases a maximum of three, and frequently
only two or one, additive componental terms is involved in the equations. It is there
fore hard to see what good may be done in the theories of elasticity or plasticity by
introducing the rules of repeated subscripts, for disregarding the writing of one or at
maximum two more existing terms in a sum, etc., but instead making it compulsory
to use new abstract symbols (such as the Kronecker delta, whose value is supposed to
represent zero for so many of the permutations of its sub- and superscripts, whereas
in the old practice no virtue needed to be made of the fact when a term was non
existent). In other words, it is hard to see what practical advantages may result
from these tendencies to upset a satisfactory perfect terminology which has conveyed
easily understandable information toengineersforgenerations. All this, too, is in view of
the commendable and unselfishly supported efforts of the three main American socie
ties of engineers in endeavoring to induce the authors of papers dealing with engineer
ing problems to accept the established standards for symbols and ways of denoting
the directions of the components of stress and strain, as proposed several years ago
and recommended since then in the "Letter Symbols for Mechanics of Solid Bodies,"
ASA Z 10.3, 1948, by the three societies.
Part II
THE YIELDING OF SOLIDS, PARTICULARLY
OF THE METALS UNDER SIMPLE STATES
OF STRESS
CHAPTER 15

LIMITING STATES OF STRESS IN SOLIDS.


THEORIES OF MECHANICAL STRENGTH
15-1. Criteria of Failure. One of the most important problems in
strength of materials is to determine the mechanical conditions which
cause solids in engineering structures to deform permanently or to frac
ture. Ordinarily it is assumed that there are a variety of states of stress
in which solid bodies may either greatly change their shapes or else fail
by fracture. In the consideration of the danger of a failure occurring
in an engineering structure several criteria must be distinguished. It
was pointed out in Chap. 3 that the permanent or plastic deformation in
ductile metals when the stresses increase may develop either suddenly
or quite gradually depending upon whether a definite yield stress charac
terizes the metal or no definite stress exists at which it starts to deform
permanently and that an observed yield point depends also on the previ
ous stressing and plastic deformation of the material. Apart from similar
conditions, the temperature is of determining influence on the magnitude
of the forces which are required to deform solid bodies.
Some of the limiting conditions upon which the design of machine parts
must be based are:
1. The load or loads under which the first permanent distortion begins
to develop if the material has a sharply defined yield stress.
2. The stresses or loads under which the permanent portions of the
strains will not exceed certain small more or less arbitrarily chosen limit
ing strain values.
3. The maximum permissible elastic displacement or deflection in a
part of a construction or in the whole construction.
4. A load under which the elastic equilibrium of the stresses in a con
struction or in parts thereof may become unstable causing elastic buckling
or sudden collapse.
5. Similarly instability of the elastic and plastic equilibrium (this
may include cases in which a load causing permanent distortion reaches
an analytic maximum after which a local deformation develops).
6. The load causing fracture, including failure through fatigue frac
ture under oscillating loads.
7. Resonance in the vibrations in a construction or parts thereof
under oscillating loads.
175
176 THEORY OF FLOW AND FRACTURE OF SOLIDS

At elevated temperatures:
8. The maximum
permissible permanent deflection or displacement
during the expected service time due to long-time creep in metal parts.
9. The maximum permissible load which is sustained continuously in
a construction part during the expected life of it without causing fracture.

At very low temperatures:


10. In certain metals which are known to be ductile the danger that a
sudden brittle fracture may occur, particularly under impact conditions,
when the load increases very rapidly and when sharp notches (concen
trators of stress) are present. The design of machines operating at high
pressures and speeds, at normal or low temperatures, raises some of
these problems.
From the preceding remarks we see that several of the criteria on which
the limiting values of the loads must be based depend on the time or on
the velocities with which the small permanent strains may form. In
spite of these facts we shall at first disregard in the following treatment
of the mechanical conditions causing sudden yielding or fracture the time
and the rates of the permanent strains as variables and shall assume that
plastic yielding as well as fracture in solids depends solely on the state of
stress or strain or on a combination of both. Examples of the excluded
cases will be treated later.
15-2. Limiting Surface of Yielding. A state of stress in a material
element is determined by six quantities, e.g., the three principal stresses
and the directions of the principal axes. A state of stress which is just
necessary to produce failure by plastic yielding or by fracture in an ele
ment may be described by the three principal stresses <r1,a%,a3. These
three quantities may be represented by the rectangular coordinates of a
point P. The totality of the points P representing different states of
stress just necessary to produce sudden yielding or plastic deformation
in a given material forms a surface

= 0
fi(<fi,<r,,<r2) (15-1)

which we may call the limiting surface of yielding of the material. It will
be one of the first problems of the mechanics of the plastic states of
crystalline solids to determine from observations the shape of this surface
for various materials.
Most solid materials will withstand very high hydrostatic pressures
without fracture if the pressure acts uniformly from all sides as it does in a
fluid surrounding the solid. Materials with a loose or porous structure,
such as wood, will undergo considerable permanent deformation under
high hydrostatic pressure and will remain condensed in their volume after
LIMITING STATES OF MECHANICAL STRENGTH 111

the pressure is released. (Wood when sufficiently condensed in this


manner will no longer float in water.) The crystalline solids (metals,
impervious consolidated natural rocks) under these conditions, however,
are compressed chiefly in an elastic way by very small amounts. With
respect to their compressibility the impervious polycrystalline or amor
phous solids behave like liquids. These are elastically compressible
bodies and will withstand high hydrostatic pressures to almost any
possible value without suffering a permanent distortion after the pressure
is released. In less compact solid materials subjected to fluid pressure,
however, marked evidences of failure have repeatedly been observed,
e.g., in marble cylinders which were exposed to hydrostatic pressure
(Karman) and in wood, which is compressed into irregular shapes
because of its cellular anisotropic structure (A. Foppl). If certain pre
cautions are not taken when such materials are exposed to high fluid pres
sures, the liquid used to exert stress may penetrate the material through
the fine fissures or cracks which it contains. It has been observed by
T. C. Poulter that glass balls which were exposed for a short time to very
high fluid pressure usually did not break at the maximum of the pressure
but either during the period of decrease of the pressure or later after it
was rapidly released. The small amounts of fluid which may have been
able to penetrate through the invisible fine surface cracks into the outer
most layers of the balls cannot escape fast enough from the crevices
when the pressure is quickly released. A gradient of the pressure in the
fluid which is entrapped in the narrow channels may thus be produced,
with the result that the surrounding material is stressed locally by tensile
stresses of high concentration when the external pressure is released,
which tends to burst or explode the glass. The crumbling through the
external pressure of a fluid can destroy the structure of weaker materials,
such as marble and sandstone. This can be prevented, as Karman has
shown, if the test specimen is covered by a very thin and flexible foil of
brass or copper which does not permit the penetration of fluid into the
crevices of the material.1
From these observations we draw the conclusion that a pure state of
hydrostatic pressure o\ = <n = <n = —p(p > 0) cannot produce plastic
1 Further noteworthy observations on this bursting action of a liquid used to trans
mit pressure were made by Bridgman, who found that cylinders of hardened chrome-
nickel steel were less able to withstand an internal pressure if the liquid transmitting
the pressure was mercury instead of viscous oil. It appears that the hardening cracks
in this material aid in the premature destruction of the pressure vessel, if the small
atoms of the mercury are able to penetrate these cracks. On the other hand, the
large molecules of the oil are not able to penetrate the cracks so easily. The pres
ence of stresses set up by quenching may have a further deleterious effect upon the
strength of vessels exposed to high hydrostatic pressures. In this connection see
also Sec. 15-12.
178 THEORY OF FLOW AND FRACTURE OF SOLIDS

distortion in compact crystalline or amorphous solid material, but only


a small reversible elastic contraction, if the pressure fluid is prevented
from penetration into the invisible crevices of the solid. The limiting
surface of yielding [Eq. (15-1)] cannot intersect the straight line

on its negative side.


15-3. Limiting Surface of Rupture. In contrast to this behavior under
high pressure it is certain that solids under a uniform tension acting in
all directions are able to resist only certain definite stresses. If the three
principal stresses are equal tensile stresses solid materials break without
preceding permanent deformation. This is even the case in many solid
materials under a simple uniaxial tension or when two of the three
principal stresses are positive and different from zero, which suffice to
rupture a body. The so-called brittle materials (glass, the cast metals,
most natural rocks) under uniaxial or multiaxial tension break suddenly
and do not deform permanently appreciably under such stresses before
they fracture. We are thus led to the hypothesis that in the region of
tensile stresses there corresponds a totality of limiting states of stress
(?i,(ri,<T3 capable of rupturing a body which may be represented by a second

surface
M<ri,<n,<n) = 0
(15-2)

which we may call the limiting surface of rupture.


15-4. Triaxial Tension. The determination of the conditions of frac
ture in solids from experiments is a much more difficult task than that
of the conditions under which materials become plastic. For various
physical reasons, some of which will be mentioned below, it is there
fore not yet possible to establish the precise form or the forms of the
surface of rupture, although some plausible assumptions may be helpful
for predicting the probable shape or shapes of it. The most important
states of stress causing brittle types of fracture known to occur under
certain conditions in otherwise ductile metals are those states in which
at least one principal stress is of tensile nature. Perhaps the most
interesting critical example of these states is that of equal triaxial ten
sions. One of the first difficulties to be mentioned which are encoun
tered in the experiments on brittle fracture is that it was not yet possible
to produce this critical state of stress in a satisfactory manner.

A few ingenious ways have been suggested. The Russian physicists A. F. Joffe
and J. Frenkel (Joffe and Levitskaia, J. Phys. Chem. U.S.S.R., vol. 58, p. 45, 1926)
proposed first to undercool a sphere of material (rock salt) to a temperature which is
much lower than the ambient temperature of the atmosphere and subsequently to
warm up the sphere in the air until it reaches room temperature. The higher surface
LIMITING STATES OF MECHANICAL STRENGTH 179

temperature of the sphere causes it to expand. The thermal stresses in the sphere
consist of compression stresses in circumferential direction in the outer portions of
the sphere, and the equilibrium of the stresses requires that the central portion of the
sphere is stressed in tension. At the center of the sphere thus a state of equal triaxial
tensions is produced. It is not difficult to compute the distribution of these thermal
stresses in a sphere of elastic material during this period of thermal heat exchange
which would depend on the radially symmetric distribution of the temperature, a
problenYwhich has been solved in the classical theory of the conduction of heat through
a sphere. The maximum values of the thermal tensile stresses in the center of the
sphere were computed by Frenkel, and it was claimed that in rock salt which had
been undercooled in liquid air they would reach high values which have never been
observed in simple uniaxial tension or bending tests of this material (the spheres of
rock salt upon rewarming did not show any cracks). These very high values of the
triaxial tensile strength of a weak material such as rock salt at an internal point of a
body must be questioned. The outer portions of the rock-salt sphere which were
stressed essentially under biaxial compression must certainly have yielded plastically
because of the low limit of plasticity this material possesses. Since the high values of
the tensile stresses were computed by using the theory of elasticity, the effect of yield
ing in the outer portions of the sphere, which tends to reduce the compression stresses
in the outer shell, was not considered, and consequently the tensile stresses in the
central portion were overestimated to a considerable degree.
M. J. Manjoine at the Westinghouse Research Laboratories proposed another
related method in 1944 ("Development and Testing of Triaxial Tension Specimens,"
Publications of the Welding Research Council, New York, Vol. 9, p. 643, December,
1944) by which he wished to utilize the volume expansion which accompanies the
gamma-alpha transformation in iron. According to H. Scott (Proc. ASTM, vol. 13,
pp. 829-847, 1928) this transformation in certain iron-nickel alloys similar to invar
can be depressed below room temperature. The test specimen in which triaxial equal
tension is to be produced is a copper sphere embedded in a hollow sphere of the special
nickel-iron alloy, to which it is soldered by copper brazing after the hole has been
closed by a plug through which it was inserted. When this composite copper and
iron-nickel alloy sphere is slowly cooled below the critical temperature, the outer
shell will contract first but at the transformation temperature it will suddenly increase
its volume. This will stress the copper sphere in triaxial tension. Preliminary tests
indicated promising results, but quantitative strength values of the copper sphere are
not yet available.
Some other methods for creating equal tensions in three perpendicular directions
were suggested by M. Hetenyi (loading of a bolt having a cylindrical head by uniaxial
tension. Photoelastic tests in the plane problem of the bolt head have shown that a
"neutral point" exists at the center in the bolt in which the shearing stresses vanish) ;
by H. C. Boardman (a cube of material pulled across each face by tensile stresses) ;
by A. P. Young, J. Marin, and others (loc. cit.). In the literature on elastic concen
tration of stresses the statement has been made that a body of revolution having a
groove in circumferential direction approaching the profile of a sharply bent hyperbola
is stressed in a state of triaxial tension in the central portions of the minimum cross
section when the body is loaded in tension in the axial direction (Kuntze, W., Koha-
sionsfestigkeit, Mitt. devt. Malcrialprufungsanstalt., Sonderheft 20, 1932. See also
Z. Physik, 1931, p. 32). The exact solution for the case of the deep hyperbolic notch
in an elastic body of revolution subjected to tension in the direction parallel to the
axis has been worked out by H. Neuber in his book "Kerbspannungslehre," Fig. 47,
180 THEORY OF FLOW AND FRACTURE OF SOLIDS

p. 80 (Verlag Julius Springer, Berlin, 1937, 163 pp.). This book is available in an
English translation, prepared by the David Taylor Model Basin, Navy Department,
Washington, D.C., under the title "Theory of Notch Stresses" (Edwards Bros., Inc.,
Ann Arbor, Mich., 1946). This solution shows that the maximum tensile stress in
the axial direction acts at the inner circumference of the notch. This stress for deep
notches is several times larger than the peripheral stress and the stresses at the center
of the bar. Thus deeply notched specimens of ductile metals must start to yield
first along the circumference of the minimum cross section before they will break, and
the stresses at fracture cannot be computed from their values derived from the theory
of elasticity (MacGbegor, C. W., and author, Proc. ASTM, vol. 34, Part II, p. 216,
1934).
L. H. Donnell, of the Illinois Institute of Technology, Chicago, 111., proposed
(Minutes Weld Stress Committee meeting, New York, Oct. 19, 1945) an interesting
way in which he hopes to prevent this premature yielding in a body of revolution
with a deep hyperbolic notch by means of eight narrow slots cut by a saw in four
equally spaced radial planes intersecting each other in the axis of the test specimen.
These slots do not reach the minimum section but are near enough to it. Their
purpose is to subdivide the bar in eight sectors in each of which additional bending
moments are artificially to be introduced which bring down the axial stress in the
outer portion of the minimum section of the grooved bar to such low values compared
with the stresses present in the central region that the yielding is prevented and frac
ture in triaxial tension in the center may be expected.

Much thought has been given by D. J. McAdam1 to the determination


of a plausible shape of the limiting surface of rupture Ji{o-\,<ri,ai) = 0 for
the metals from long series of his experimental investigations. Although
he made use of the representation of this surface in three-dimensional
plots (<7i,<r2,<r3) he preferred in many cases also to limit the discussion to
the states of stress characterized by two equal principal stresses, which
could be represented in a plane figure by plotting the two equal stresses
ai = «r3 as abscissas and the stress <r\ as ordinates. In connection with
the subject of this section we shall limit reference to his work to the
representation of his test results for the brittle metals in this (<ri,<rs = <rs)
plane diagram. He distinguishes in it two limiting curves a and b
(Fig. 15-1). His fracture stresses were observed either in simple tension
(point A) or simple compression (point B) or in tension tests with super
posed lateral hydrostatic pressure (branch AC of curve a) or by rupturing
round bars having a deep circumferential notch of varying proportions
in tension tests. (The latter tests furnished the points in the branch .4 D
1 McAdam, D. J., Jr., G. M. Geil, and W. H. Jenkins, "Influence of Plastic
Extension and Compression on the Fracture Stress of Metals," ASTM 1947 preprint;
McAdam, G. W. Geil, and It. W. Mebs, Influence of Plastic Deformation, Com
bined Stresses and Low Temperatures on the Breaking Stress of Ferritic Steels.
Metals Technol., August, 1947; McAdam, "Fracture of Metals under Combined
Stresses," ASM 1945 preprint; The Technical Cohesive Strength of Metals in Terms
of Principal Stresses, Metals Technol., December, 1944; and previous papers.
LIMITING STATES OF MECHANICAL STRENGTH 181

of curve o and branch DE


of curve b.) In these tensile tests with notched
bars ffi thus represents the axial stress and o-2 = 03 the radial stresses at
the axis of the bar in the minimum section of the notch. Test point C
refers to a test with a round bar subjected to a hydrostatic pressure on the
cylindrical surface, without stress in
AXIAL STRESS
axial direction, and test point D to the
state of equal triaxial tension stresses
ffl = o-j = o-j.1

If a brittlematerial should break


under any combination of stresses as
soon as at least one of the three prin
cipal stresses is a tensile stress and
when this stress reaches a certain max RADIAL STRESS

imum value independently from the


value of the two other principal stresses,
the test points in Fig. 15-1 would lie
along the two dotted lines c and d . The
lines c and d represent a theory of the
Fia. 15-1. Limiting states of stress
fracture stresses (maximum stress the for fracture of brittle metals accord
ing to D. J. McAdam.
ory) which will be mentioned in Sec.
15-13a. Thus we note that, according to the experiments of McAdam,
the latter theory of fracture is not verified, since a metal breaks under
a state of pure triaxial equal tensions
<Tl
= 02 = ffi > 0 ,

(point D in Fig. 15-1) at a stress which is nearly twice as large as the


strength in simple tension (<n > 0, <r2 = <r3 = 0, point A). It is worth
noting that in McAdam's fracture diagram (Fig. 15-1) a single continuous
curve (such as a or b) represents all the states of stress causing fractures
when all three principal stresses are positive (tensile) stresses (the
portions of the curves a and b in the positive quadrant 0-1,0-2) and when all
three stresses are negative (the portions of curves a and b in the opposite
quadrant of a 0^,02 system of coordinates). It is known that in brittle
materials the types of the fractures which are observed under these two
conditions are different in character (they are of the cleavage and shear
types). It may be remarked that observations in tests with polycrystal-
line brittle materials (such as cast iron, marble) to which reference will
be made (see Chap. 20 on the compression test) indicate that under uni

1 From the preceding, we note incidentally that McAdam's diagram (Fig.


15-1) is
the orthogonal projection on the a\,az (or o-l,<r3) coordinate plane of a plane section
through the surface of rupture fz(ai,a^,ai) = 0. This plane passes through the a\
axis and contains the space diagonal <n = <n = «■,.
182 THEORY OF FLOW AND FRACTURE OF SOLIDS

or biaxial compression stresses the oblique shear fractures predominate,


which may follow other rules than the cleavage-type fractures which run
perpendicularly to the largest tensile principal stress. It also seems
probable that in Fig. 15-1 the distances OA and OE should be equal. A
state of equal biaxial tensile stresses (point E, ax = 0, <r2 = a3 positive)
can easily be produced in a thin-walled hollow sphere or cylinder of a
brittle material (cast iron or glass) by internal pressure. In such burst
ing tests probably the same fracture stress would be observed as in simple
tension. Many valuable results of his tests on the influence of previous
cold working on the increase of the fracture stresses in ductile metals, low
temperatures, etc., in their effects on the shape of fi(ai,<ri,<r3) = 0 cannot
be reported here, and the reader is referred to his papers.
It will be mentioned later (Sec. 15-13/) that, according to A. Leon, some
of these conclusions from McAdam's tests concerning the existence of a
single "limiting curve" representing the states of stress causing both the
cleavage and the shear types of fractures in brittle materials may be
reconciled by the Mohr theory of strength, to which reference will be
made below.
16-5. Cleavage and Shear Fracture. By comparing the appearance
of the broken surfaces of specimens, two types of fractures have been dis
tinguished. Brittle materials such as glass or cast metals when tested in
prismatic or cylindrical specimens in tension break in a plane which is
perpendicular to the direction of the tension. In amorphous materials
the surface of fracture appears smooth, but such materials sometimes
show a "conchoidal" fracture with minute concentric ripples when the
tensile stresses causing the fracture have been produced through the blow
of a concentrated impact force. In crystalline materials, on the other
hand, a rupture surface caused by simple tension, when this surface
appears to be perpendicular to the direction of tension, consists of minute
planes in each of the crystallites not coinciding with the plane of fracture,
which reflect the light like small mirrors. This kind of rupture surface
ispopularly called "crystalline" fracture. In axial compression, how
ever, prismatic specimens of most solids break in a surface which is
inclined at a certain angle, usually smaller than 45 deg, with respect to
the direction of compression (cf. Fig. 20-3). Ductile metals when tested
in tension break after having contracted permanently locally near the
minimum section in both manners: the central portion of the fracture
surface appears perpendicular to the direction of the tensile force, the
outer portions of it in round bars form a cone inclined at an angle of
approximately 45 deg with respect to the axis of the specimen ("cup and
cone" fracture; Fig. 15-2). Specimens which have been broken in axial
compression also frequently show a fragmented surface, parts of which
are covered with a fine dust of detached particles while other portions
LIMITING STATES OF MECHANICAL STRENGTH 183

of the fracture surface appear polished. The former type of rupture


along a surface which is perpendicular to the direction of tension has been
called cohesive or brittle failure, the latter shear fracture. There seems
to be a fundamental difference in these two types of fracture, which is
better understood if one observes the manner in which large crystals of
material break. The mineralogists, when desiring to obtain samples of

Fig. 15-2. Low-carbon steel bar broken in tension. Original diameter of bar 3 in.

crystals or rock salt or calcite from pieces in nature, etc., wedge a sharp
knife into them and produce their specimens by "cleavage." If stresses
of a tensile nature are applied by an impact, crystals of an ionic or homo-
polar lattice break in clean-cut surfaces along a crystallographic plane
(the cubic crystals of rock salt, for example, along cubic planes).1

1Careful observation has shown that an apparent "cubic cleavage" plane really
consists of numerous parallel cleavage planes and of many minute steps between them
(Kusnetzow, Z. Physik, vol. 42, p. 302, 1927; Tertsch, Z. Krist., vol. 78, p. 53, 1931 ;
vol. 85, p. 17, 1933; vol. 87, p. 326, 1934). The opinion has been expressed that rock
salt under slight impacts cleaves in many small steps simultaneously more easily than
along one single mathematic crystallographic plane.
.

184 THEORY OF FLOW AND FRACTURE OF SOLIDS

E. Schmid and W. Boas1 have carefully studied the phenomena of


fracture in crystals. According to them, small specimens of rock salt
consisting of one crystal tested in tension break when the tensile normal
stress in one of the cubic planes nearest to the plane perpendicular to the
tension direction reaches a constant value: "the cubic tensile or cohesive
strength." A similar condition may hold at very low temperatures in
certain metallic single crystals. Zinc crystals belonging to the hexagonal
system, according to Boas and Schmid, break along a base plane after
little preceding plastic slip, if these crystallographic planes are favorably
oriented relative to the tension direction.2 The metals of the cubic
systems, however, behave in a complicated manner at normal tempera
tures because of the numerous opportunities their crystals offer for slip
in their lattices. Single crystals of these metals are extremely ductile
and may be stretched considerably by simple slip until double slip may
start in them, causing a maximum of the tensile force. After this is
reached, they start to neck. When they finally break, the crystal ordi
narily has contracted to a thin band and a fracture plane may not be
easily distinguishable. It is presumed that the "fibrous" appearance
of the broken surfaces of pieces after shear fractures occurred in them in
the polycrystalline cubic ductile metals may be due to this complex
behavior of their single crystals.
C. F. Tipper3 distinguishes three ways in which the fracture takes place
in the crystals of aggregates under tensile stresses: (1) the fracture plane
coincides with a plane of slip (shear) in the crystals, (2) a wedge is formed
by two shear planes symmetrically oriented to the direction of the major
principal tensile stress, and (3) by cleavage. Many variations are pos
sible, however, in the details which constitute the facets of the broken
surface. In mild steel she found two types of fractures, one in which the
crystals cracked after little deformation and the other in which they
1 "Kristallplastizitat," pp. 132, 175, Verlag Julius Springer, Berlin, 1935, 373 pp.
* It may be added that no satisfactory mechanism has yet been proposed explaining
what happens along the atom or molecule rows when a crystal breaks by a cleavage
fracture. As Orowan ("The Solid State of Matter," p. 81, International Conference
on Physics, London, 1934, Vol. II, 183 pp.) pointed out, one could logically postulate
either that a cleavage fracture starts simultaneously from all atoms or molecules in
the cleavage plane or physically assume that it starts at a certain point or a few points
and runs through the lattice of the crystal with a finite velocity. Most investigators
have accepted the second hypothesis, assuming that a finite time is required for the
formation and the propagation of a cleavage crack. Around the leading edge of the
crack a very high elastic concentration of the stresses probably develops which pene
trates with a large velocity as a wave into the lattice which is accompanied by such
violent oscillations of large amplitudes of the atoms or elements of the lattice that the
bonds between the latter are hroken.
» The Fracture of Metals, M etallurgia, vol. 39, p. 133, 1949.
»

LIMITING STATES OF MECHANICAL STRENGTH 185

were drawn out through plastic distortion, giving the broken surface a
"fibrous" appearance.
We have mentioned the "cup and cone" fracture of round bars of
ductile metals in the tension test. While the conical portion of the
fracture surface is interpreted as a shear fracture,1 there is disagreement
among observers as to whether the central portion of the fracture surface
is of the cleavage or the shear type. In longitudinal sections through the
axis of the bars one sometimes sees the central portion of the fracture
surface as a zigzag line in which each step is inclined at an angle of approx
imately 45 deg with respect to the tension direction. J. B. Friedman2

(o) (b) (c)


Fig. 15-3. A magnified idealized
section normal to a fracture surface. The crystallites
a and b were split by cleavage along cubic planes; c broke along shear planes; d is a view on
the shear fracture planes along which the crystallite c broke.

believes that the central portion of the "cup and cone" fracture consists
of shear fractures. E. R. Parker, Harmer E. Davis, and A. E. Flanigan,3
however, found that the ferrite crystals of mild steel bars, when broken
in tension at very low temperature, showed cleavage fractures; at inter
mediate temperatures the central portion of the cup and cone fracture
was a combination of cleavage and shear fractures; and at normal temper
ature the shear planes predominated in the ferrite grains. This was
clearly shown by them by etching several plane sections through the
rupture surface and by observing the relative orientation of the small
etching pits which had formed in the crystals under a microscope (see
Fig. 15-3). These pits in ferrite crystals consist either of small three-
sided pyramids having an equilateral triangle for their base (when situ-
1 O. Mohr (Joe. cil.) proposed to explain fracture along the cone by means of his
theory of strength as a shear fracture.
* About the Mechanism of Failure of Metals, /. Tech. Phys. U.S.S.R., vol. 11,

No. 11, 1941 (translated by Westinghouse Research Laboratories and the David
Taylor Model Basin, Navy Department, 1947). See also a monograph by the same
author, "Mechanical Properties of Metals," p. 122, Institute of Materials of the
Aviation Industry, Moscow, 1946, 423 pp. (Russian).
2 A Study of the Tension Test, Proc. ASTM, vol. 46, p. 1159, 1946.
186 THEORY OF FLOW AND FRACTURE OF SOLIDS

ated in an octahedral plane of a crystallite) or of minute inverted four-


sided pyramids; (the four sides of their base disclose the directions that
are parallel to the sides of the elementary cube of the body-centered
lattice in the ferrite grains).1 It is thus established that what may appear
as a "cleavage fracture" may consist predominantly either of planes in
the crystal grains oriented in the directions in which the ferrite grains
deform plastically by slip (shear planes with respect to the crystallites)
or of planes coinciding with the true
cleavage planes (cubic planes) of the
ferrite grains or of a combination of
both kinds of (slip and cleavage)
planes, depending on the tempera
ture and the rate of loading. Since
the central portion of the tensile cup
and cone fracture in round bars of
ductile metals after very fast rates
of stretching or at low temperatures
is made up of cleavage planes in the
Fio. 15-4. Cleavage fracture in cubic crystal grains, we note that poly-
planes of a very pure iron broken by ten crystalline metals can break in this
sile impact at room temperature. [Ac
knowledgments for the micrograph (.magnifi manner after they are deformed perma
cation 70) are due to Colonel G. F. Jenks nently to some extent. The cleavage
and H. C. Mann of Watertown Arsenal
(1937). Iron "Puron" (99.95% Fe, fracture usually associated with a
0.0,5% C, 0.040% Oi) made by Dr. T. D. "brittle" behavior is not confined
Yensen, Westinghouse Research Labor-
torics.] only to the cases in which no appreci
able permanent deformation occurs
but may be observed in metals that have been permanently deformed
before fracture.2
The transition from ductile to brittle fracture in plates of steel of the
types used in ship hulls has been studied in static tension tests with wide
flat specimens of the full thickness and in appreciable sizes at the David
Taylor Model Basin of the Navy Department. The method consisted
in cutting an internal slot perpendicular to the tension axis in the plates.
To obtain different sharpnesses of the two internal notches, small holes
were drilled at the ends of the slot or fine saw cuts made there with a
jeweler's saw (Fig. 16-5). In this manner it was possible to produce
1 A typical cleavage fracture in iron coinciding with the cubic planes of the crystals
which was obtained in a high-speed tensile test is illustrated in the micrograph in
Fig. 15-4.
• Further observations on the shape of the fractured surfaces of round bars pulled

in tension and simultaneously subjected to a high lateral hydrostatic pressure are


reported on p. 271.
LIMITING STATES OF MECHANICAL STRENGT 187

brittle (cleavage) fractures in large flat specimens of steel tested in static


tension at normal temperatures.1
16-6. Influence of Rate of Loading, Time at Load, Stored Energy in System, and
Size of Specimens on Fracture Stress. If a flat strip of glass is supported horizontally
at both ends and loaded transversely in its center by
a weight, it has been observed when the load is left
acting over some time that the rod may suddenly
break after having carried the load even for a pro
longed time. This has been observed particularly in
strips cut from a glass plate by a diamond. The sharp
notches left along the edges of the strip by this crude
method of splitting glass may contribute to the effect.
It appears that it takes a certain time for a cleavage
crack to progress after it starts from one of the fine
notches situated in one of the edges in the region of
maximum tensile stresses near the section in which
the maximum bending moment acts. Similar obser
vations have been made on glass or porcelain bars hav Fig. 15-5. Tensile specimen
ing a ground or the original fused surface, however. having internal notches used
by David Taylor Model
Two groups of tests made at the suggestion of the Basin.
author may further illustrate this. Round bars of
porcelain of the shape shown in Fig. 16-6 with a ground surface were tested in
tension by increasing the load continuously at different rates. Since a fired porcelain
is essentially an elastic material, in which stress and elastic strain increase pro
portionally, these tests may be considered as constant-rate-of-load as well as
rate-of-elastic-strain tests. The results are reproduced in Fig. 15-7.2 The relative
velocities between the two heads of the testing machine in inches
per second are shown on a logarithmic scale as abscissas, the rup
ture stresses as ordinates over a range of 10s in. /sec. The specific
rates of elastic strain in the gauge length of the porcelain bars
were approximately one-twentieth of the relative head rate. The
tensile strength of the porcelain was found to increase as the log
arithm of the rate of strain. A few additional tests made on an
impact machine especially designed for very rapid tension tests for
3% brittle materials furnished values (not shown in Fig. 15-7) on the
extension of the straight line. The times required to break the
specimens in these very fast tensile tests were of the order of
0.0001 sec. From these tests it must be concluded that a brittle
material, under ordinary conditions when tested in air, does not
possess a definite tensile strength but that the latter continuously
decreases with the time during which a load can act. Similar
Fio. 15-6. results were observed in so-called "time load" tests, in which
glazed or unglazed porcelain round bars were held in a rack loaded
by a constant bending moment over periods of several months. From the high-speed
tensile tests it would appear also that, contrary to the usual assumption, porcelain is
stronger under impact conditions than when the load is applied slowly.
1 Windenbttrg, D. F., Notch Toughness of Steel Plates, Product Eng., vol. 19,
p. 110, 1948.
* From an unpublished report by R. K. Carlson, June,
1941.
188 THEORY OF FLOW AND FRACTURE OF SOLIDS

The wide range of static tensile strength in various glasses and in porcelain was
carefully investigated by F. W. Preston and his collaborators.1 They found that the
tensile strength of these materials depends on the rate of application of the stress or
on the time a constant load is carried on a specimen. In a long-time test the lower the
load is chosen the longer it will be carried. When they represented their test results
by plotting the reciprocal values of the stress above the logarithms of the duration of
stress, they obtained straight lines for a variety of the investigated amorphous

*
8*

10'* 10'' 10 *
MEAD RATE OF TESTING MACHINE M/sec.
Fio. 16-7. The effect of the rate of loading on the rupture stress of porcelain in tension.

materials (Figs. 15-8, 15-9, and 15-10). The long-time tensile strength <r of these

materials may thus be expressed by

--G) (15-3)

(<r„ a constant stress, t the time the constant stress a acts, U a constant of the dimension
of a time).1
In a quick-loading device they were able to apply a force in the center of a rod sup
ported at both ends on knife-edges, stressing it in bending in 2 to 3 millisec, maintain
it for 10 millisec, and then remove it in % millisec. In additional devices they could
vary the times of loading and the duration of the maximum load over a wide range.
Special precautions were taken to keep the temperature and the moisture content in the
air surrounding the test rods under control. They found that the strength depended
greatly on the moisture content in the air, even in silica glasses. Ordinary glasses
were found 20 per cent stronger when dry than when wet and 2 to 2.5 times as strong
after drying and when tested in vacuum in 10 sec as compared with short-time tests
made in air at atmospheric pressure. The chemical attack of water on glass hereto
fore overlooked under atmospheric temperatures was even more pronounced at ele-

1 Three papers by F. W. Preston and T. C. Baker and one by Preston and J. L.


Glathard in the same issue of the </. Applied Phys., vol. 17, No. 3, pp. 162-195,
March, 1946.
* These results were further discussed in a remarkable paper by Nelson W. Taylor,
who proposes a theory of the breaking strength and of the cleavage fracture in amor
phous substances based on the ideas of H. Eyring laid down in the latter's theory of
rate processes, including the effect of temperature (see p. 200).
LIMITING STATES OF MECHANICAL STRENGTH 189

vated temperatures and pressures. These tests disclose the great importance of
controlling more precise conditions in the surface of such brittle substances as glasses
(or porcelains) in any strength experimentation (the films of moisture which are
deposited on the surface of the test specimens influencing greatly the values of the
breaking stresses) quite apart from the significance of the duration of loads affecting
the observed strength values. According to these authors, thoroughly dried glass in
high vacuum appears to have a tensile strength fairly independent of the time of

I I I I I L_ I I I
O.I I 1, 1,* I,3 I,4 I,S o.l I 10 I02 I03 10* 1,s
DURATION OF STRESS IN SECONDS DURATION OF STRESS IN SECONDS
Fio. 15-8. Annealed soda-lime glass, Flo. 15-10. Porcelain, type B, dry.
wet.
Fios. 15-8, 15-9, 151-1,. Fracture stresses for glass, silica, porcelain. {Testa by GUUhart
and Preston.)

loading for short durations of load, and glasses exposed to moisture rapidly lose their
strength under stress. This may decay to one-third of its value in vacuum in 24 hr.1

A few observations in bursting tests which were made by E. A. Davis2


with hollow cylinders of a crystalline material, a medium-carbon steel
under combined stress will illustrate the influence of the energy stored in
1 Thin films of fluid substances brought upon the surface of solid bodies may cause

a variety of the most complex phenomena when the solids are put under stress. Fluids
present near the surface of solids may exert a weakening or a strengthening effect due
to chemical actions or physical influences or both (corrosion, capillary action, enhanc
ing of the plastic deformability of a substance, e.g., by water surrounding a stressed
rock-salt crystal, embrittlement of a surface layer, the strengthening of a brittle
porcelain through a well-fitting glaze after firing, etc.). Concerning some of these
complex phenomena cf. the paper by C. Benedicks, The Effect of Wetting, Rev. mit.,
pp. 9-18, January-February, 1948.
*J. Applied Mechanics, September, 1948, pp. 216-221. For further details on
cleavage and shear fractures of low-carbon steels see also Sec. 17-7.
190 THEORY OF FLOW AND FRACTURE OF SOLIDS

the pressure fluid (oil) stressing a specimen on the type of fracture observed
in steel. Hollow cylinders with closed or open ends were stressed by
internal fluid pressure. In one group of the tests the cylinders were
connected with a large tube of good steel which served essentially only
as a tank for storing additional large amounts of energy in the oil used as
pressure fluid. The second group of the specimens was tested without,
being connected with the large cylinder. On the plastic behavior of the
test specimens the difference in energy could have no effect, and in all
cases the rupture started as a short shear fracture first in axial direction
along a plane inclined at an angle of 45 deg with respect to the surface
of the tube and parallel to its axis. After the oil started to flow out
through the shear crack that had formed, however, a pronounced differ
ence in the behavior of the specimens could be observed. When the test
specimens were connected with the pressure tank, the velocity with which
the crack spread greatly increased to such high values that the shear fail
ure changed into a cleavage crack which was normal to the surface of the
tube. In tne tests in which only low energies were available, the fracture
remained of the shear type. One "herringbone" (cleavage) fracture is
reproduced in Fig. 15-11, and one of the test cylinders after rupture is
shown in Fig. 15-12. The tests clearly indicate that the type of fracture,
whether shear or cleavage kind, depends also on the speed with which a crack
is forced to spread. In a few cases some of the cracks, which first started
slowly and later developed into "herringbone" frac
as shear fractures
tures, again in their last portions became shear fractures after their speed
of growing slowed down.
It seems, though, from all these observations that one fundamental
type of fracture of crystalline solids is the "cleavage fracture" caused by
separation in the grains of polycrystalline material along certain crystal-
lographic planes when the tensile normal stress in these planes reaches a
definite mean value, particularly in the lower range of temperatures and
at the higher rates of stressing — in contrast to what is popularly desig
nated as the "shear fracture," the conditions of which have been still
less well observed and are little understood.

The complex nature of the fracture phenomena may be illustrated by a final remark.
If cylindrical bars of a hrittle material such as fired porcelain are tested in tension or
bending and the size of the specimens is varied, it has been claimed by observers that
their tensile strength decreases with increasing size of the test pieces supposed to be of
the same grade of material. Similar observations have been reported by comparing
the tensile strength of geometrically similar round bars of different size which were
all machined from the same original stock of mild steel.1 Whether a pure effect of
1 Davidenkov, E. Siievandin, and F. Wittmann (Leningrad), J. Applied Mechanic*,
ASME December meeting, 1946.
LIMITING STATES OF MECHANICAL STRENGTH 191

the size of geometrically similar specimens on their tensile or bending strength exists
in materials which were strained only elastieally before fracture is still an open ques
tion, because of the extreme difficulty in ascertaining homogeneity in the test samples
in various sizes (e.g., in such a material as fired porcelain). The same is true for sam
ples cut from mild steel or another ductile metal that has been previously rolled cold

Fio. 15-11. Fig. 15-12.


Fio. 15-11. Cleavage(herringbone) fracture in hollow cylinder of steel.
Fio. 15-12. Hollow cylinder of steel broken by internal pressure (high-stored-energy test).
In the center of the axial crack the short shear fracture can be seen on the right side along
which the tube burst first. The remaining portions of the crack are cleavage fractures.
(Tests by E. A. Davis.)

or hot or forged. Postulating the possibility of the existence of a "size effect" on


the fracture stresses in brittle materials such as fired porcelain, W. Weibull1 advanced
a statistical theory of the strength of materials, to which reference must be made here
explaining the lowering of the strength through the greater probability of fractures
occurring in larger than in smaller specimens. An influence of the space gradient of
stress on strength in specimens tested in pure bending or torsion, which has also been
suspected in materials, is a phenomenon of a similar nature.
1 Proc. Royal Swedish Institute for Engineering Research (Ingeniors Vetenskaps
Akademien), Stockholm, No. 149, 27 pp.; No. 151, 45 pp.; No. 153, 55 pp.
192 THEORY OF FLOW AND FRACTURE OF SOLIDS

16-7. Velocity of Propagation of a Cleavage Fracture. M. Greenfield and G. Hud


son1 have measured the velocity with which a cleavage fracture runs through poly-
crystalline steel which was tested in a state of simple tension. The flat bar specimen
of steel contained a very fine saw cut as a notch on one side which initiated the cleav
age crack in the bar. The velocity of its advancing edge was measured by recording
the times at which fine wires carrying electric current which crossed the path of
the crack were broken. They found that a cleavage crack advances in steel with a
velocity of 1,000 m/sec = 1 km/sec, which is about one-fifth of the velocity with
which an elastic shear wave travels in the same material. In contrast to the fast

Fio. 15-13. Fia. 15-14.


Fios, 15-13 and 15-14. Spiral cracks in safety glass disks broken under a concentrated
load.

rates with which cleavage cracks develop, direct visual observation indicates that
shear fractures require comparatively longer times to form ; on large steel plates tested
in tension one may see them opening in times of the order of several seconds, depend
ing on the rates at which the heads of the testing machine are driven. Round bars
develop the shear crack in the cup and cone fracture rather rapidly in the ordinary
tensile tests under the rates applicable in tensile machines.
In Figs. 15-13 and 15-14 patterns of cleavage and spiral cracks in round disks of
double laminated safety glass, which were observed by L. H. Hettinger,' are repro
duced. The disks of 0.135 in. thickness and 2 in. diameter were simply supported
along their periphery and loaded by a concentrated force at the center or excen-
trically. The majority of the radial cracks crossed only the layer a of the glass on
the opposite side to the side 6 of the composite disk which carried the concentrated
load, although a number of the radial cracks could also be seen penetrating through
both glass layers a and 6. The most interesting feature in the patterns of cracks is

1 Proc.Natl. Acad. Sri., vol. 31, p. 150, 1945.


• Acknowledgments are due to L. H. Hettinger and to Willson Products, Inc.,
Reading, Pa., for having kindly supplied the test specimens showing these fractures.
See also a paper by Hettinger in Natl. Glass Industry, January, 1942, and ASTM
Bull. 147, August, 1947, p. 13, for a photograph of one of these fractured plates.
LIMITING STATES OF MECHANICAL STRENGTH 193

the crack which runs in circumferential direction almost as a perfect Archimedes


spiral having a constant pitch through the layer a of glass. The spiral did not pene
trate layer 6. Since the tangential bending stress <rt in a simply supported circular
elastic plate loaded by a concentric force at the center is larger than the radial bending
stress a, at every point except at the center where both stresses are equal, it is certain
that the radial system of cracks must have formed first in layer a. Because of the
presence of the thin layer of glue binding the plates a and 6 no glass fell off, and the
numerous segments in which layer a was subdivided were held together. After the
tangential stresses <r, in the layer a had vanished, the segments could carry only radial
stresses. Although their distribution is not clear because of the friction forces which
would prevent the further motion of the segments relative to each other and because
of the glue layer transmitting shearing forces in the radial direction in a composite
plate, it may be noted that conditions apparently developed under which the spiral
crack could form. These examples may illustrate the finite time required for the
formation of the spiral crack in the circular glass plates and perhaps also effects of
flexural elastic oscillations which influenced the radial pattern of rupture.
The propagation of cracks in glass was observed by E. F. Poncelet1 by using the
high-speed motion-picture equipment of H. E. Edgerton* at the Massachusetts Insti
tute of Technology. The former discussed several conditions suspected of influencing
the speed with which the leading tip of a crack advances in a stressed plate or body of
glass, while the latter measured this velocity, which he found varying between wide
limits, as one would expect because of the dynamic circumstances influencing the
distribution of the stresses during the process of the fracture. Edgerton found crack
velocities in glass of the order 1 to 70 m/sec, Schardin3 as high as 1,500 m/sec. As in
the formation of the slip planes in single metal crystals seen on microscopic scale and
in the development of the plasticized layers in certain polycrystalline metals on a
macroscopic scale, we may note that the beautiful regular pattern of radial cracks or
of concentric ripples (in the conchoidal fractures) of glass broken under a concentrated
force does not form at once but successively. It is presumed that in a glass plate bent
by a concentrated force one radial crack forms first. This releases the stresses in the
direction perpendicular to it. Because of the inertia forces accompanying the rapid
formation of this crack, the redistributed tangential bending stresses must have their
maximum values along the line which bisects the polar angle of 180 deg, and the second
radial crack will form along this line. After the plate is split in four equal quadrants
around the point at which the force acts, the circumferential bending stresses develop
their new maximum values along the four lines bisecting the right polar angles, and
the plate splits along the latter lines in eight segments. This process may repeat
itself again and again, with the final result seen in the regular distribution of radial
cracks in Figs. 15-13 and 15-14.

16-8. Fracture along Grain Boundaries. It has been observed that


the metals in tension under a continuously acting load at high tempera
tures develop a kind of brittle fracture. The rupture surface, however,
runs along the grain boundaries perpendicular to the direction of tension
1 "Fracture and Comminution of Brittle Solids"; Metals Technol., AIME, April
1944.
■ Barstow, F. E., and H. E. Edgerton, Glass Fracture Velocity, J. Am. Ceram.
Soc., vol. 22, p. 302, 1940.
3 Schardin, H. and W. Struth, Glaslech.
Ber., vol. 16, p. 219, 1938.
194 THEORY OF FLOW AND FRACTURE OF SOLIDS

and the structure frequently shows deteriorations (Figs. 15-15 and 15-16)
along many similarly located surfaces ("intercrystalline" fractures in
contrast to the "transcrystalline" cleavage failures which were just
mentioned, occurring at low temperatures).1
From the preceding examples it thus appears that fracture is the out

Wthe
growth of various processes taking place in
grain structure, some of which require
a considerable time, others a short time to
develop, and some of which as in the case
of creep fractures weaken the grain bound
aries and may be connected with chemical
changes (oxidation of metal or of one of the
constituents of an alloy along the grain
boundaries).2

15-9. Stress Concentration. Effect of Notches


and Flaws. It is well known that sharp reentrant
corners and notches in stressed bodies cause frac
tures through the concentration of the stresses
around the fillets. Their presence is particularly
dangerous in the brittle amorphous materials
stressed elastically, but they may also cause pre
mature fractures in the ductile polycrystalline
metals under rapidly increasing loads or at low7
temperatures when the stresses are of a tensile
nature. It is perhaps less well known that certain
singularities in the stress field may initiate the
shear fractures in materials under compression
stresses. A singularity in the stress field can de
velop at the corners of prismatic specimens or
Fig. 15-15. Flo. 15-16. along the edges of cylinders compressed in axial
direction between rigid compression plates; when
Flos. 15-15 and 15-16. Fractures
along the grain boundaries. Ten the latter prevent the lateral spreading of the
sile tests at high temperatures by compressed material because of the frictional
M. J. Manjoine. forces exerted by the compression plates, the sin
gularity in the stress field arises because the
shearing stresses must vanish in the free sides of the prisms or in the surface of a
cylinder, near the compression plates, while they may have large finite values in

1 Bad "forgeability" of metals is probably due to a similar cause, weak grain


boundaries.
1 In
this connection special reference is made to the work by J. Neil Greenwood and
his associates at the University of Melbourne in Australia on the intercrystalline
fracture and the creep of pure lead and its alloys. Cf. Greenwood, J. N., and R. S.
Russel, The Influence of Impurities on the Properties of Lead, Proc. Australasian
Inst. Mining & Met.; new series, No. 87, September, pp. 135, 145, 167. See also
Bleakney, H. H., reprint ASTM 1947 meeting, containing a review on intercrystal
line fracture phenomena.
LIMITING STATES OF MECHANICAL STRENGTH 195

these planes themselves. [The condition t»» = ryl or of the equality of the com
ponents of shearing stress in two perpendicular plane sections valid in an internal
point of a stressed body (Chap. 9) is evidently not satisfied at the sharp corners them
selves, if the shearing stress has a finite value in the compression plane.1] In narrow

\\\\\\\'

Fio. 15-17. Fig. 15-18. Fio. 15-19.


Flos. 15-17, 15-18, and 15-19. Singularity in stress field causing initiation of a shear
fracture in compression test.

wedge-shaped regions (Figs. 15-17 and 15-19), this causes either a severe destruction of
the structure in brittle materials (rocks) or a concentrated yielding under large shears
through which a shear fracture is produced in ductile materials. (Cf. the views of
tension or compression specimens in Figs. 18-27 to 18-30 and several examples in
Chap. 20.) This localized yielding and slip which was initiated through very small
notches in rock-salt crystals was studied by A. W. Stepanow* in Joffe's laboratory by
1 A distribution of plane stress satisfying this condition has been described by the
author (Z. Physik, vol. 30, p. 115, 1924) where also the angle under which a shear
fracture starts at the corner was computed.
1 Physik. Z.
Sowjetunion, vol. 8, p. 25, 1935; vol. 12, pp. 183, 191, 1937. Also J.
Phys., Acad. Sciences, V.S.S.R., vol. 3, p. 421, 1940. Interesting samples of a frac
tured plastic (organic material) were shown the author
in 1945 by Drs. Contorova, Frenkcl, and Stepanow in
Leningrad. The specimens broken in tension tests showed
numerous minute steps in the cleavage plane of the shape
of ordinary hyperbolas (Fig. 15-20). Suppose that the
cleavage fracture starts from many small centers of stress
concentration (flaws). As the tensile stress reaches a cer
tain value from each center, a plane crack will develop
perpendicular to the tension direction and will spread
radially with a finite velocity. Suppose that two centers a
short distance apart are situated in two parallel planes Fio. 15-20. The steps
between cleavage planes
close to each other and perpendicular to the direction of
form branches of hyper
tension and that two plane cracks along these planes start bolas.
with a small time lag from each center to grow. The
geometric locus or curve along which these two plane cracks approach nearest each
other must be a hyperbola, since it is the locus of the points having a constant dif
ference of the distances from two fixed points (the focuses of the hyperbola). Along
this curve a step will form connecting the two cracks. Figure 15-20 shows these
steps along the branches of these hyperbolas as they appeared in actual broken
samples.
196 THEORY OF FLOW AND FRACTURE OF SOLIDS

means of the photoelastic method in connection with the formation of cleavage


fractures.
Severely cold drawn or rolled metals show a fibrous grain structure. The grains
are elongated and the nonmctallic inclusions are drawn out along streaks in them, thus
weakening the strength of the metal considerably if it is stressed by tensile stresses in
the direction which is perpendicular to the direction of rolling. This may be the
case even in hot-rolled metals in impact tests, as Davis1 has shown by testing small
prismatic specimens having a sharp V notch (izod impacts amples) in their central
section in impact bending tests. These specimens were cut from a hot-rolled steel
round bar either in longitudinal or in radial direction. The plane through the notch
in them coincided either with the plane of a cross section or with a plane passing
through the axis of the round bar. The specimens cut in the latter direction broke
at impact energies which were 25 per cent
smaller than the energies required to
break the other specimens.
Interesting observations on the con-
ehoidal type of cleavage fractures were

EThe
made by C. A. Zapffe, F. K. Landgraf,
and C. O. Worden in their studies on
transgranular fractures in metals.5

15-10. Fracture Theory of Griffith.


significance of minute faults
«?WiP *~f"i$J0*'- S3 (flaws) or of small invisible cracks
Flo. 15-21. Conchoidal cleavage fractures assumed to be present in most solids
observed in cast molybdenum. (Magnifi
cation 180.) (Zappfc, Landgraf, Worden.) has been emphasized by Griffith3 in
his theory of the brittle fracture of
amorphous materials (glass, fused silica) and demonstrated in his frac
ture tests with fused silica bars. He showed that very thin filaments of
glass or silica which had been drawn above 730°C when immediately
tested at room temperature had a surprisingly high tensile strength, which
increased with decreasing diameters:

Filament diameter, mm 1.02 0.0508 0.0178 0.0033


Observed tensile strength, kg/cm' 1,750 5,610 11,500 34,600

1 J.
Applied Mechanics, March, 1945.
■ Metal Progress, September, 1948, p. 328. In broken pieces of molybdenum they
found fine radiating ripple marks and undulations in the fracture surface within many
grains in the polycrystalline structure. A micrograph of such a fractured specimen
is reproduced in Fig. 15-21. In these observations the authors used a technique
developed by C. A. Zapffe and M. Clogo, Jr. (" Fractography — A New Tool for
Metallurgical Research," preprint for 26th annual meeting of the ASM, 1944) for
studying the untouched cleavage facets of broken crystal aggregates of metals at high
magnifications.
3 Griffith, A. A., The
Phenomena of Rupture and Flow in Solids, Phil. Trans..
vol. 221, pp. 163-198, 1921.
LIMITING STATES OF MECHANICAL STRENGTH 197

The high values of the tensile strength are due to the capillary action
of the surface tension in very thin filaments. The surface tension carries
a proportionally increasing fraction of the load with decreasing diameters.
This condition prevails for a short time only, and the internal stresses
generated in these materials after cooling and the injuries received after
exposure to the atmosphere (dust, humidity, etc.) very soon are sufficient
to decrease the high strength values to the ones which are ordinarily
observed.
Griffith has furthermore investigated the condition under which a small
crack inside an elastic body must spread when the material surrounding
the crack is stressed by a constant externally applied system of forces.
Suppose that the crack has just formed. New surface energy has thus
been produced. Simultaneously, because of the formation of the crack
the stresses around it must redistribute themselves. They will be
substantially reduced near the flat portions of the surface and will be
very sharply increased around the edge of the crack. The result of the
formation of an elongated cavity, however, will be a total decrease in the
potential energy of the internal elastic forces. Griffith solved two cases.
In one of these he assumed that, in a thin strip of elastic material stressed
in tension, an elliptic hole would form having its major semiaxis a oriented
in perpendicular direction to the direction of tension a\. The maximum
stress at the apex of the major axis of an elliptic hole has been computed
by Inglis1

(16-3o)

b being the minor semiaxis. The radius of curvature at the apex of the
ellipse being p = 6'/a, it follows that the maximum stress due to concen
tration of the stresses in a very flat ellipse or in a narrow crack of length
2c having a very small "notch" radius (fillet radius) p is approximately
given by
<rmn, = 2<ri (15-4)

if the 1 in the former formula is neglected. Griffith evaluated the amount


of energy (referred to the unit of length of the thickness) by which the
(potential) elastic energy decreases when an elliptic hole of length 2c is
formed and found it equal to

W. =
^ (15-5)

(E modulus of elasticity). If T denotes the surface tension in kilograms


1 Trans. Inst. Naval Architects London, vol. 60, p. 219, 1913.
198 THEORY OF FLOW AM) FRACTVRE OF SOLIDS

the newly formed energy of surface tension per unit of thickness of the
strip is
W. = 4cT (15-6)
and the energy has therefore decreased by

W = W. - W. =
E
- 4cT . (15-7)

The elliptic crack will increase its length if the energy W becomes an
analytic minimum (or if — W, representing the increase in energy or the
total energy, becomes a maximum) or if dW/dc = 0, which gives
I2ET
(15-8)

Thus the tensile strength of a strip of glass in which short cracks of


length 2c are present in a direction perpendicular to the stress <ri should be
inversely proportional to the square root of c. The corresponding
formula for two equal tensions acting in perpendicular directions was
tested and verified by Griffith in bursting tests with thin-walled spherical
shells of glass in which fine cracks were made by a diamond.
To explain the considerably higher strength of amorphous brittle
materials in the ordinary compression test Griffith ingeniously postulated
that large tensile stresses exist around the tips of the small elongated flaws
in a material under uniaxial compression if the flaws are inclined at a
certain angle with respect to the direction of compression. Because of
the concentration of the stresses, however, the tensile strength around
the flaws is reached under a much higher mean compression stress (eight
times as high) than in a field of simple tension. The Griffith theory
therefore explains in a plausible manner mechanically the well-known
fact that the compression strength in brittle materials is much higher than
the tensile strength, although the predicted value of the ratio of both
may seldom be observed. E. Poncelet1 remarks that this discrepancy
might be due to spontaneous formation of new flaws while the compres
sion stress increases, an explanation not yet considered by Griffith.
The high tensile strength of very thin filaments of fused silica and of
various glasses has also been determined by F. O. Anderegg.2 For very
thin flame-blown silica fibers which were loaded stepwise, the load being
released after each short duration, he found:

3.1 4.1 6.6 8 4


Tensile strength, kg/cm' 367,000 174,000 121,000 97,000

1 Lor. cit.

•Strength of Glass Fiber, Ind. Eng. Chem., vol. 31, pp. 290 -298, 1939.
LIMITING STATES OF MECHANICAL STRENGTH 199

Fracture as an Instability of Equilibrium.


15-11. The great value
and importance of the Griffith theory are in having shown that the tensile
strength 9\ of amorphous solids depends on a few physical variables or
parameters (the "average" length 2c of the hypothetical flaws or cracks,
the modulus of elasticity E, and the surface tension T) and that the con
ditions of fracture in the cases treated by him can be understood by
investigating the equilibrium of the stresses around these hypothetical
weakened regions of small but finite size (large compared with the atomic
distances), in which an exchange of energies of different character lakes place
causing the propagation of growth of the weakened regions.
The observations recorded above should point to the fact that the
phenomena of fracture might be discussed either by determining the
mechanism of separation within the crystal lattices of solids or by investi
gating whether the mechanical equilibrium of the forces in stressed regions
of finite dimensions might become unstable because one type of energy
may change to another form during the process of weakening of a struc
ture, which precedes the final rupture.

It cannot be the purpose of this book to develop in detail the means based on the
atomic theory of solids for computing the tensile or cohesive strength of solids, and
reference in this respect must be made to books on physics dealing with the atomic
structure of matter.1 It has been stated (Chap. 7) that the values of the tensile
strength of single crystals of the metallic elements, of their alloys, and of chemical
compounds (e.g., of rock-salt crystals) which were computed from the ideal lattice
forces were found to be 100 to 1,000 times larger than their observed values.' Expla
nations for the causes of this discrepancy between the theoretical and the observed
values of the cleavage strength and of the shearing stresses required for plastically
deforming a crystal by slip have been proposed.

1 Seitz, Frederick, "The Modern Theory of Solids," 1940, 698 pp. and "The
Physics of Metals," 1943, 330 pp.; Barrett, C. S., "Structure of Metals, Crystallo-
graphic Methods, Principles, and Data," 1943, 566 pp.; Glasstone, S., K. J. Laidler,
and Henry Eyring, "The Theory of Kate Processes," 1941, 611 pp.; all published
by McGraw-Hill Book Company, Inc., New York. Mott, N. F., and H. Jones,
"The Theory of the Properties of Metals and Alloys," Oxford University Press, New
York, 1936.
2 For example, by M. Polanyi
(Z. Physik, vol. 7, p. 323, 1921), who equated the
work done for separating two atom layers in a crystal lattice by their distance against
their attractive forces to the work newly created by the surface tension; or by F.
Zwicky (Physik. Z., vol. 24, p. 131, 1923) for the strength of sodium chloride (rock-
salt) crystals. These theories and the imperfections that have been proposed for
explaining the discrepancy between the atomic and the actual strength in various
solids have been reviewed in "Papers and Discussions of the International Conference
on Physics," London, 1934, Vol. II
"The Solid State of Matter," Cambridge Univer
sity Press, Ixmdon, 1935.
200 THEORY OF FLOW AND FRACTURE OF SOLIDS

Investigators have assumed that the crystals contain numerous small imperfections
of one or another kind in their lattice.1
The effect of the time at load on the cleavage strength of glass and similar brittle
solids to which reference was made (page 188) has led Nelson W. Taylor* to propose a
theory of the cleavage fracture for such materials. He regards the empirical relation
of Glathard and Preston [quoted above in Eq. (15-3), Sec. 15-6] between the fracture
stress a and the time t at load as a measure of the velocity of a molecular process which
is controlled by an activation energy just as the rates of chemical reactions are influ
enced by such an energy. Assuming that Hooke's law holds until a bar of glass breaks,
that the strongest chemical bond in the substance can be stretched a certain charac
teristic length in order to be broken, which might be expressed as the elastic extension
of a corresponding length, depending on the stress a and the modulus of elasticity E,

1 Those postulated by L. Prandtl and by Sir Geoffrey Taylor have been mentioned
in Chap. 7. Further ingenious forms for the imperfections have been proposed by
A. Smekal (Z. tech. Phyrik, No. 11, 1926; Phys. Z., vol. 27, p. 837, 1926), by A. A.
Griffith (loc. cit.), by F. Zwicky (Helv. Phys.
Ac<<lj yol 3 p 2^ 1930; proc jy^ Af0d
DIRECTION OF SLIP

vol. 15, p. 253, 1929; in the "mosaic structure"


suggested by him, consisting of a periodic dis-
tribution of weakened layers of atoms within a
crystal lattice dividing it into regular blocks),

(b)-^^^^Qg^S^2?SIi3' Some experiments by G. I. Taylor and


Fio. 16-22. Fragmentation of la- H.J. Gough (Edgar Marburg lecture, Proc.
^ CryStaI' (After J' AS™' voL 33' P- ^ 1933) disclosed exact
Got'T)
information relative to fine periodical changes
in the orientation of the slip planes of aluminum single crystals after plastic
deformation. Both investigators made measurements of the intensity of the X-ray
reflections from the slip planes in specimens. From such tests Gough and W. A.
Wood (Strength of Metals in the Light of Modern Physics, J.
Roy. Aeronaut. Soc.,
August, 1936, p. 17) concluded that (1) the crystal lattice is slightly distorted along
the direction of slip while it is practically undamaged in the transverse direction and
(2) the X-ray evidence might be due either to very shallow waves in extremely thin
lamellae elastically bent or more probably (3) these fine lamellae are broken up into
minute fragments or blocks, as shown in Fig. 15-22 periodically tilted at very small
angles relative to the plane of slip of the ideal crystal. Concerning these various
types of periodic imperfections cf. also the books quoted on p. 199 and R. Houwink,
"Elasticity, Plasticity and Structure of Matter," p. 102, Cambridge University Press.
London, 1937.
A divergence of opinions has been expressed concerning the question whether the
imperfections should be assumed in the interior of the material or only localized essen
tially in the surface of stressed specimens. Smekal, Zwicky, and Griffith supported
the former, Joffe and Orowan the latter view. Orowan showed that the tensile
strength of thin mica sheets is considerably increased if the tensile stresses are relieved
along the free edges of the sheets; he achieved this by applying two equal tensile
forces in opposite directions by means of suitable clamps attached to the central por
tions of a sheet of mica (loc. cit.; London Conference on the Solid State, 1934, p. 89:
and Z. Physik, vol. 82, p. 235, 1933).
*J. Applied Phys., vol. 18, p. 943, 1947.
LIMITING STATES OF MECHANICAL STRENGTH 201

and after introducing an activation energy for rearrangement of the atomic structure
to permit the extension of the strongest bond until it is broken, Taylor finds for the
time / at load the formula:
t = taeKa/'kT (15-8a)

(U, k, a constants, T the absolute temperature). After taking logarithms an equation


results:

ln-t

ft)
On ,
=
a to (" (15-8fc)

This was the relation found from rupture tests with glass [Eq. (15-3)] described in
Sec. 15-6 which expressed the fracture stress a as a function of the time at load.

I
It to be hoped, on the other hand, that further attempts will also be undertaken
is

to define satisfactory mechanisms of the second kind, which were just mentioned, for
explaining the various modifications of cleavage, shear, and mixed fractures which
have already been observed, including the effect of previous cold working in the
ductile metals on the type of rupture.1

We have mentioned the reorientation of the grain boundaries in parallel direction


1

and the alignment of nonmetallic inclusions along the direction of rolling in deformed
metals, and we may include in the same category of effects cases in which material

a
in ita unstrained condition contains a large number of small elongated zones of weak
ness of plate shape (flaws, cracks). In a quasi-isotropic material these flat flaws must
be statistically oriented, in equal numbers for each inclination of their middle planes
(Fig. 15-23), but after the material deformed, the flaws change their relative orienta
is is

tion in the body. the strain predominant in one direction and of sufficient
If

Fio. 15-23. Preferred orientation of small cracks (flaws) after a considerable strain.

magnitude, a proportionally larger number of them will be found in positions in which


their middle planes are inclined at a small angle with respect to the direction of the
major principal strain than at larger angles. The question whether the tensile
strength of the deformed material in the direction parallel to the major principal
strain increased or not much changed requires further careful consideration. The
is

is

tensile strength of the deformed material in the perpendicular directions, however,


will certainly be found to be smaller.
Miss Cantorova (according to an oral communication, 1045) explained in this

f
202 THEORY OF FLOW AND FRACTURE OF SOLIDS

It has been mentioned that in polycrystalline solids two types of


rupture have been distinguished by observers: the cleavage and the shear
fracture. The most regular and simplest prototypes of the crystalline
solid metals, namely, the large metallic single crystals, also show these two
types of fractures; they have their characteristic cleavage planes distinct
from their shear planes. Under impact or when certain conditions pre
vail (very high strain rates or low temperatures) they break along the
former, after considerable plastic distortion by slip (or twinning) in thin
layers along the latter planes when their capacity for this distortion has
been exhausted. In the polycrystalline ductile metals after moderately
large permanent strains and under homogeneous states of stress, the two
types of fracture are superficially characterized through the orientation
of the fracture surface; if it is perpendicular to the largest tensile principal
stress, it is usually called a cleavage, if inclined at a considerable angle
with respect to the principal directions of stress, a shear fracture. It has
already been pointed out, however, that a microscopic check of the sur
faces of separation may disclose important differences; an apparent
cleavage fracture may consist of minute true cleavage planes or of minute
shear planes (in the individual grains). It may even become necessary
to establish the percentage in area of the first against that of the second
type of characteristic plane surfaces for deciding whether the fracture
consists predominantly of cleavage or of shear fractures. This leads to
some paradoxical conclusions. To these may be added observations
that the grain structure may be broken up along the grain boundaries.
One or several of these elementary three types of processes of separation
may contribute to the fracture.

manner the dependence of the tensile strength on the direction in the strained condi
tion of materials. The effects of cold work on raising or lowering the tensile strength
of deformed ductile metals for different directions may depend on some such mechan
ism ; an important factor to be considered, however, must also be whether the preced
ing deformation was produced by lateral compression stresses, as in the case of cold
drawing or rolling, or by pure tensile forces parallel to the direction of stretching.
The well-known "slaty cleavage" found in samples of natural rocks which have been
severely deformed by simple shear during geologic times of increased mobility (moun
tain folding periods) in the strata of the earth is another example of variable strength
caused by a prior deformation in a material.
A theory predicting fracture stresses in materials which have either only deformed
elastically prior to rupture or also permanently has been advanced by E. Saibel.
(Metals Technol., Nos. 2131, 2186, 1947). It is based on thermodynamic concepts
and on the assumption that the elastic or plastic strain energy used up in a metal
before the fracture occurs by tensile stresses is equal to the mechanical equivalent
of the heat of fusion times the change in volume of the substance on passing from
the solid to the liquid state divided by the volume in the solid state at the melting
temperature.
LIMITING STATES OF MECHANICAL STRENGTH 203

To elaborate further on this point, suppose that torsion tests are made
with specimens having the shape shown in a longitudinal section in the
sketch (Fig. 15-24) by twisting the two ends around the axis of the hollow
specimen by two moments. If a ductile material is used, maximum slip
in the minimum section will develop for an increment of distortion along
the directions shown by short lines in a side view (a) of the specimen ; and,
since the fracture coincides with the direction of one family of these lines,
in which the maximum shearing stresses act, this will be described as a
shear fracture. (We may incidentally note, on the other hand, that the
other family of maximum shear planes which is parallel to the axis of the

Fio. 15-24.

bar must behave in a very inactive manner, since they project into por
tions of the specimen in which the stresses decrease rapidly. Thus, if the
bar breaks along one of the latter shear planes, the crack stops soon
and has no tendency to grow.) Suppose now on the contrary that a
brittle material is tested. Round bars of such materials are known to
break in torsion tests along a helix perpendicularly to the direction of the
tensile principal stress inclined at an angle of 45 deg with respect to the
axis. Suppose that the material has a little ductility. This is soon
exhausted because the tensile strength is reached in those inclined direc
tions which are marked in the side view (b). In a perfectly homogeneous
material one expects that a large number of these short inclined tensile
cracks will form and, if a sufficient number of them has formed, that the
bar will break. During the fracture probably most of the small project
ing saw teeth will be sheared off and the fracture surface may disclose
the characteristics of a shear fracture. A microscopic check may show
the minute inclined tension cracks. Is this a "shear" or a "cleavage"
fracture? According to the orientation of the elementary small tension
cracks we believe it should logically be called a tensile failure. Yet the
macroscopic fracture surface runs parallel to the direction of the maximum
shearing stress!
204 THEORY OF FLOW AND FRACTURE OF SOLIDS

Careful observation of the system of surfaces in which prismatic speci


mens of marble or similar brittle rocks break in ordinary compression tests
shows that the macroscopic rupture surfaces frequently follow the direc
tions of minute surfaces of slip in certain portions of the compressed
prisms, while in other portions they are of the cleavage type. L. Prandtl
and F. Rinne described such observations to the author many years ago
(see the sketch in Fig. 15-25). Although admittedly the distribution of
the stresses in these prisms of compression tests is far from being uniform
for well-known reasons (see Chap. 20), these cases will illustrate the

Fiu. 15-25. Fio. 15-26. Fig. 15-27.


Fig. 15-25. Complex shear and cleavage fracture.
Fig. 15-26. A crack consisting of cleavage, shear, and grain boundary failures.
Fig. 15-27. Steps between surfaces of slip forming a fracture surface.

important fact that even in those portions of compression specimens in


which a simple homogeneous state of stress is present the fracture surface
may select directions which cannot be simply referred to the principal
axes of stress (see two more examples in the sketches, Fig. 15-26 and
15-27). If the external load is smaller for a complex fracture consisting of
heterogeneous elements (cleavage, shear, grain boundary cracks, etc.), this
will develop. In the future it will perhaps be possible to analyze these
conditions by statistical means, if the elementary processes are better and
quantitatively determined, contributing to the various causes of the
separation in the grain structure of a solid material under the different
types of stress. Thus the problem of fracture becomes a determination
of smallest loads sufficient to cause a possible instability of the static
equilibrium of stresses in a complex grain structure with which trans
formations of different types of energy into each other may be associated,
or the localized formation of minute steps of cracks or of plastic slip in a
macroscopic layer which leads to a failure. In spite of a quasi-isotropic
solidly joined grain structure, subjected to a perfectly homogeneous state
of stress, finally, internal singularities of the field of stresses may develop
LIMITING STATES OF MECHANICAL STRENGTH 205

from internal flaws or regions of weakness, concentrating the causes for


further spreading of a failure.
16-12. Fracture of Heavy-walled Hollow Cylinders of a Brittle Material under
Internal Pressure. The results of a few fracture tests under combined stresses with a
brittle material may be of interest.1 Heavy-walled cylinders of porcelain were loaded
under internal pressure until fracture occurred. In one series of tests short cylinders
of 5 in. inside and 8 in. outside diameter with a cylindrical gauge length of 1 1 in. having
conically enlarged heads wrere loaded by internal hydrostatic pressure. The ends

Fio. 15-28. Fio. 15-29.


Fios. 15-28 and 15-29. Test cylinders of porcelain.

were cemented to flanged steel casings to which circular plates of steel were attached
(Fig. 15-28); the tightness against pressure was achieved by flat rubber gaskets,
which were inserted between the flat ends of the porcelain specimens and the steel
plates. The fracture stresses were computed from the Lame formulas

r2* + r,* pr,2


— —
r22 ri2 r22 ri2

(p .„ pressure; a,,at,at = radial, tangential, and axial stress at the inner surface;
ri = inner, r2 = outer radius).
An unglazed cylinder (Fig. 15-28) when water was used as pressure fluid broke at a
pressure p = 758 lb/in.2 and a tangential stress at = 1,730 lb/in.2. A similar cylinder
with the inside and outside surface glazed broke at p — 825 lb /in. 2 and <rt = 1,880
lb /in.2 Through a small alteration of the shape of the two conically reinforced
1 The test results are taken from two unpublished reports by J. L.
Lambert, M. J.
Manjoine, and W. E. Trumpler of the Westinghouse Research Laboratories, East
Pittsburgh, Pa. (July, 1940, and March, 1941).
206 THEORY OF FLOW AND FRACTURE OF SOLIDS

heads of the test cylinder (leaving the cone to project out a small amount from the
steel casing, uncovered by cement) the bursting stress of the glazed cylinder increased
to <r, — 3,335 lb/in.2 when water was used and in another cylinder to at = 4,130 lb/in.2
when mineral oil was used as pressure fluid. In the preceding tests the pressure was
increased at a rate of 20 atm = 284 lb/in.2. In a subsequent test with another
cylinder (both surfaces glazed and tested by using water as pressure fluid) the pres
sure was held for 15 min at 994 and at 1,136 lb/in.2. The cylinder shattered at the
stress a, = 2,762 lb/in.2.
To determine this noticeable influence of the nature of the pressure fluid on the
fracture stress two additional series of tests were run with smaller tubes of the same
unglazed porcelain (inside diameter 2 in., outside diameter 3J4 in.), one in which the
hollow specimens were tested with both ends open, using the arrangement shown in
Fig. 15-29, the other with closed ends. The observed bursting pressures and stresses
were as follows:

Average breaking Average breaking


Pressure fluid used
pressure, lb/in.2 stress, lb /in.2

Specimens with open ends

Nitrogen . . 2,038 4,586


Water. . . . 1,974 4,440
Mineral oil 2,315 5,210

Specimens with closed ends

Nitrogen 1,575 3,542


Water 1,756 3,955
Oil 1,918 4,320

Noting that the kinematic viscosity of a mineral oil is about 200 times that of water
at room temperature and that both are practically independent of the pressure, but
that the kinematic viscosity of nitrogen varies inversely as the pressure, being about
15 times that of water at atmospheric pressure, but only one-tenth that of water at
an average pressure of 2,000 lb/in.2, at the latter pressure the kinematic viscosities of
the pressure mediums for nitrogen 0.00106, for water 0.01, and for oil 2 cm2/sec, we
see that the bursting stresses trt, with one exception in the last column of the table,
appear in increasing order with the increase of the kinematic viscosities of the three
pressure fluids that were used.
From these bursting tests with a brittle material such as fired porcelain, we may
note that: (1) its tensile strength when the tensile stresses in hollow specimens are
produced by internal hydrostatic pressure increases notably with the kinematic
viscosity of the pressure medium ; (2) short cylinders with closed ends break at lower
pressures than the ones with open ends because the constraint of the end connections
raises the stresses locally near the conical ends: (3) glazed surfaces (i.e., surfaces which
during the firing passed through a fluid state) in general increase the strength of porce
lain against tensile stresses, provided that the glazes "fit" well, i.e., maintain the
skin near the surface in compression; and (4) the strength of a brittle material in
general decreases with the time under load.
LIMITING STATES OF MECHANICAL STRENGTH 207

Although the tensile strength in a brittle material is altered if a fluid surrounds its
surface in which a high hydrostatic pressure is maintained (because the invisible
fissures in the surface of brittle materials permit the pressure fluid to penetrate in
the cracks, causing local concentrations of the stresses), E. Morsch1 found that the
tensile strength of specimens of concrete was practically independent of the value of
moderately large compression stress which acted in one of the directions perpendicular
to the direction of tension. In other words, according to Morsch, concrete under a
state of plane stress in which one principal stress was a tensile and the second principal
stress a compression stress (the third principal stress being equal to zero) broke under
the same value of the tensile stress independently of the value of a compression stress
which acted simultaneously in one of the directions perpendicular to the direction of
tension. Further details concerning these facts supporting the strength theory of
Mohr of certain brittle materials will be reported in the following section.

15-13. Theories of Strength. As long as the mechanisms of the several


types of fractures in various materials mentioned in the preceding sec
tions have not been better determined by experiments, nothing else
remains than to formulate the conditions of fracture on phenomenological
grounds. We shall distinguish the cleavage and the shear fracture.
These two types of rupture may or may not correspond to two classes of
materials of construction which engineers describe as "brittle" and as
"ductile" materials, depending on whether they do not deform appreci
ably or deform permanently before rupture.
Tests of solids carried out under states of stress in which all three
principal stresses are negative and compression stresses (such states of
stress may be obtained after superposing a state of pure hydrostatic pres
sure of sufficient intensity on any of the ordinary states of stress, such as
simple uniaxial tension or compression, or pure shear) show conclusively,
however, that the so-called "brittle" materials without exception may,
under suitable conditions, be brought into the plastic state before a
fracture may occur. It is therefore more correct to speak, not of brittle
or ductile materials, but rather of the "brittle" or the "plastic" states of
these materials. This may be expressed also by stating that in general for
a given material both limiting surfaces fof yielding and of fracture Eqs.
(15-1) and (15-2)] have an importance in the applications.
It has been known from the early times of material testing that the
behavior of various materials under similar conditions, e.g., under simple
tension, under compression, or under torsion varies considerably. This
led the older investigators to propose several criteria for the conditions of
failure, although they did not yet possess clear concepts as to whether
these criteria should refer to failures due to flow or to fracture or to which
type of rupture they should be applied.
1 Ing. Archiv, vol. 1, p. 232, 1930. Also M6rsch, E., "Der Eisenbetonbau," 6th
ed., Vol. 1, Part 2, p. 294, Stuttgart, 1929.
208 THEORY OF FLOW AND FRACTURE OF SOLIDS

Because of the historic value of a few of the earlier theories relative to


the conditions causing failure of material by plastic yielding or by frac
ture, they are listed and briefly described here. Remarks added may
indicate cases in which one or the other of these criteria may be useful in
expressing specific conditions of failure. Following the suggestion of the
late Prof. B. P. Haigh1 and of H. M. Westergaard2 it will be instructive to
represent the corresponding limiting surfaces geometrically by plotting
them in a system of rectangular coordinates in which the coordinates are
made equal to the principal stresses a\, at, and 0-3.'
a. Maximum Stress Theory. According to this theory, the maximum
principal stress in the material determines failure regardless of what the
other principal stresses may be, provided that the latter are absolutely
smaller. This theory cannot be applied as a criterion of yielding because
three equal tensile or compression stresses as such belong to those states
of stress which should cause failure according to this theory, in contradic
tion to the experimental experience according to which such states cannot
produce a plastic (but only an elastic) distortion in compact solid mate
rial. The surface representing this theory is a cube. The faces of the
cube are six planes symmetrically spaced around the origin of the ai,aita3
system of rectangular coordinates and parallel to the coordinate planes.
Parts of this surface, however, have served as criteria of the cleavage
fracture in materials. A modified form of this hypothesis consists in
assuming a smaller limiting positive value (k,) if one of the three principal
stresses 0-1,0-2,0-3 is a tensile stress and a larger limiting negative value
|/C2|

— > ti) one of the principal stresses a compression stress.


ki,
is
if

(
The corresponding stress surface also cube; the center of no longer
is

it
a

coincides with the origin of the <r1,0-2,0-3 system of rectangular coordinates


but appears displaced along the space diagonal within the octant in which
all the principal stresses have negative values.
Maximum Elastic Strain Theory (Theory the So-called "Equivalent"
b.

of

Stress, St. Venant). This hypothesis was much used some years ago in
France and in Germany for expressing the condition of brittle fracture
and even that of yielding. According to this theory the maximum posi
tive elastic extension of the material determines failure of either kind.
If the elastic strain
- - +
=g

«i [<ri v(*2 <r,)] (15"9>


^
\

Haigh, B. P., Engineering, vol. 190, p. 158, 1920.


1

Westergaard, H. M., J. Franklin Inst., p. 627, May, 1930.


2

In review of the theories of strength by A. Mehldahl (Brown Boveri


1

Rev.,
a

Zurich, p. 260, 1944) photographs of plaster of paris models can be found illustrating-
the shapes of the surfaces of yielding according to the various hypotheses which he
compared.
LIMITING STATES OF MECHANICAL STRENGTH 209

is positive, the expression



<rl v(<r2 + <T3) i= <ro (16-10)

(where is Poisson's ratio) should be smaller than the positive limiting


v

"equivalent" stress <r0. In the case of simple uniaxial compression


(— a, a > 0) the lateral strains are positive and equal to va/E. If these
positive strains are made equal to a0/E, i.e., if a < (<r0/v) in ductile metals
a yield stress in pure compression according to this theory should be
observed equal to three times that in pure tension (assuming v = ^).
This is not the case in many metals. Although an altered improved
form of this theory has been extended to include also definite negative
limiting strains (of absolute value different from the positive limiting
value), the theory likewise cannot hold; for, under three equal principal
stresses of equal sign, either of these limiting strains could be reached and,
according to this theory, the material should fail by flow; however,
triaxial equal tensions or compression stresses do not deform a material
permanently. The fact that failure according to this hypothesis would
depend on Poisson's ratio v besides the two "equivalent" stresses for
tension and compression does not speak much in favor of it. The limit
ing surface /(<ri,<r2,<r2) = 0 corresponding to this theory consists of two
straight three-sided pyramids in inverted positions relative to each other,
having equilateral triangles as sections normal to the axis which coincides
with one of the space diagonals in the cn,<ri,<Ti system of rectangular
coordinates.
c. Theory of Constant Elastic Energy of Deformation (Beltrami). The
total elastic energy stored in a material before it reaches the plastic state
can have no significance as a limiting condition, for the same reasons,
which have been mentioned, since under high hydrostatic pressure large
amounts of elastic energy may be stored without causing either fracture
or permanent deformation.
Theory of Constant Elastic Strain Energy of Distortion (M. I. Huber,
d.
H. Hencky) or of the Constant Octahedral Shearing Stress. The strain
energy of distortion is obtained after subtracting the elastic energy
of volume dilatation from the total elastic energy stored in a material.
Using Hooke's law for elastic deformation :

(16-11)

(ti,e2,e3 principal strains, <ri,<r2,<r3 principal stress, v Poisson's ratio, E


modulus of elasticity) the total strain energy stored in a unit of volume is
210 THEORY OF FLOW AND FRACTURE OF SOLIDS

=
^ [<ri2 +<r22 + <r,2 - 2r(rm + <rv, + van)] (15-12)

If we subtract from this energy the work done which is used in changing
the volume, namely,

+ + + a + a) = - + +
(<ri <r2 <r3)(«i
^jg,2"
(at <rj <r3)2 (15-13)
^

the expression for the strain energy of distortion per unit of volume is
obtained:

W = (<Tl2 + <r22 + <T32



<T\<Ti

<r2<r3

<r3<ri) (15-14)

or

W = ~ + ~ + ~
[(<ri <r3)i ("3 (15_15)
I2G

[where G = E/2(l + p) denotes the modulus of rigidity]. For simple


tension, <ri = a0, <n = <r3 = 0, 17 = <roVoG. The condition that at the
plastic limit the elastic energy of distortion reaches a constant value
W = const may therefore also be expressed by writing

(<ri
- <r2)2 + Or2
~
T3)2 + (<r3
~ gQ2 = 2g„2 = const
(15-16)

where <r0 is a material constant.


When comparing the left side of Eq. (15-16) with the expression which
was derived in Eq. (10-27) for the octahedral shearing stress

to =
I V(*i
- <r2)2 + (<r2
- <r,)2 + (<r,
- <ri)S (15-17)

we note that the condition W = const is expressed even more simply by


stating that at the plastic limit the octahedral shearing stress t0 in the material
is constant, or that

= = const
To . (15-18)

Experiments made on the flow of ductile metals under biaxial states of


stress have shown that Eq. (15-18) expresses well the condition under
which the ductile metals
at normal temperature become plastic (see
Chap. 17). The constant
in this case is the yield stress of the metal
a0
in simple tension or compression. For simple or pure shear according
to Eq. (15-16) or (15-18) the theory of constant octahedral shearing stress
LIMITING STATES OF MECHANICAL STRENGTH 211

predicts for the principal stresses the values <t\


= — <rj = <r0/ y/Z, a 3 = 0.
For the ductile metals, Eqs. (15-16) or (15-18) therefore express the
equation of the limiting surface of yielding. This surface is a straight
circular cylinder (Fig. 15-30). Its axis coincides with the space diagonal
in the positive quadrant of the a\,ai,a3 system of rectangular coordinates.
The radius of the cylinder is equal to y/% <ro- Since this limiting surface
of yielding does not intersect the straight line which is the axis of the
cylinder, no point of the surface exists at which d = <r2 = <r3. States of
equal principal stresses are excluded from the surface, in accordance with
the experiments according to which hydrostatic pressure cannot cause
plastic deformation.
e. M aximum Shearing Stress The
ory. Observations made in the
course of extrusion tests on the flow
of soft metals through orifices led
the French engineer Tresca1 to the
assumption that the plastic state in
such metals is created when the
maximum shearing stress just
reaches the value of the inner resist
ance of the metal against shear.
The appearance of the so-called slip
layers, of fine markings on the
surface of deformed bodies, which
approximately coincide with the
Fig. 15-30. Surface of yielding for con
directions of maximum shearing stant octahedral shearing stress to =
stress, seemed to lend much support constant.
to this assumption. These fine line
markings were interpreted as the intersections of thin layers of material
with the surface of the deformed pieces, in which the grain structure of
the substance was distorted through the yielding. These planes in cer
tain materials are inclined at an angle of 45 deg with respect to the direc
tions of the largest and smallest principal stress.2 Assuming <r\ > <r2 > <r3,
the condition of yielding according to this theory is that the maximum
shearing stress reaches a constant value or that


tmi - — Pi—a3
<T\
= const . (15-19)

1
Compl. rend. Savants itrangers, 1865, 1868, and 1870, Paris.
2 The strain or flow figures which accompany the transition from the elastic to the
plastic state have been studied particularly for mild steel and will be described lates
in more detail (cf. Chaps. 18 and 37).
212 THEORY OF FLOW AND FRACTURE OF SOLIDS

The behavior of metals under triaxial unequal compression stresses and


also under other states of stress in the plastic range of strains is predicted
by this theory in a satisfactory manner, although the results of experi
ments are better represented in many cases by the theory of the constant
octahedral shearing stress. According to both these theories the yield
stress for uniaxial tension and compression should be equal to each other,
which is the case for several of the ductile metals. For simple or pure
shear <ri = — <rs = <r<>/2 for Tm., = o-0/2 = const whereas

go

for Tct = const. According to the maximum shearing stress theory the
"slip lines" should be inclined at an angle of 45 deg with respect to the
directions of the principal stresses o\ and <r3. This we find to be approxi
mately true in mild steel. On the other hand, for brittle crystalline
materials, which cannot be brought in the plastic state under uniaxial
tensile stresses, but which may yield a little under simple compression
before they break, the angle of the slip planes which may become dis
tinctly visible or of the shear fracture surfaces which usually develop
along these planes differs considerably from the planes of maximum shear.
Likewise, in these materials the values of the tensile and the compressive
yield stress if both are observable are not equal. The latter fact is obvi
ously in contradiction to the maximum shear theory, which, as we have
seen, demands that the yield stresses in uniaxial tension or compression
should be equal for a given material. The condition of flow [Eq. (15-19)]
according to this theory does not contain the "intermediate" principal
stress o-j, which can have any value between ci ^ o% ^ aj. This condition
in its most general form is expressed by three equations:

0i — ffj = ico , at — ci = +co , as — C2 = i^o (15-20)

where the absolute value of the yield stress in tension or compression.


<r0 is

Thus we note that the surface of yielding corresponding to the maximum


shearing stress theory consists of three sets of parallel planes. Each set
is perpendicular to one coordinate plane and parallel to the plane bisecting
at an angle of 45 deg the angle of the two coordinate axes which appear
in the corresponding equation of the two parallel planes. These six planes
define a straight hexagonal prism in the ai,<Tz,(n system of coordinates.
The cross sections of the prism are regular hexagons. The axis of the
prism coincides with the space diagonal of the positive quadrant of the
<T|,(rj,tr3 axes. It is not difficult to verify that this hexagonal prism is
inscribed in the straight circular cylinder representing the theory of
LIMITING STATES OF MECHANICAL STRENGTH 213

constant octahedral shearing stress t„% = const described in the previous


section. (According to both theories, certain states of stress must coin
cide in their principal stresses, e.g., <ri = <r0, <r2 = a3 = 0, - • • , etc.)
The theory of maximum shearing stress may also express the condition
under which shear fractures occur in ductile metals after prior plastic
distortion. E. A. Davis1 found, for example, that heavy-walled cylinders
of a medium-carbon steel tested under internal hydrostatic pressure and
an axial load after having deformed plastically burst in surfaces which
were inclined approximately at 45 deg with respect to the major and
minor principal stress and that the maximum shearing stress in these
surfaces had approximately a constant value at fracture.2
E. G. Thomsen and J. E. Dorn' arrived at similar conclusions by testing
rupture of magnesium alloy tubes.
If one of the principal stresses vanishes, e.g., if <r3 = 0 either the theory
of constant octahedral shearing stress or the theory of maximum shearing
stress may be represented in the same plane figure. Using the principal
stresses a\ and <r2 as rectangular coordinates, obviously such a figure
represents the plane intersections of the corresponding two surfaces of
yielding of a straight circular cylinder and of the straight hexagonal
prism which were just mentioned with the coordinate plane a\,<ri. The
former intersects the <r% = 0 plane in the "plasticity ellipse"

<ri2
- <t«t2 + <r22 = <ro2 (15-21)

1
Cf. the test results quoted in Sec. 15-6, p. 190.
*The shear fractures were oriented in two groups. When the ratio n of the true
tangential stress <r, to the true axial stress aa at fracture was 0 < n < 0.762, the follow
ing stresses were observed at fracture:

n ="
<rt/<ra 0 0.500 0.750 0.762
56,500 55,000 55,800 55,000

and the shear fracture surfaces were all cones coaxial with the axis of the hollow
cylinders. When, however,

0.775 0.800 0.875 1.000 2.000 00

Tm„, lb/in., 44,100 41,100 39,700 41,600 37,700 38,150

the shear fractures occurred in planes parallel to the axis of the specimens. In each
group the mean value of rn, was very nearly constant. The difference in these two
groups of fracture stresses was due to the anisotropy of rolled stock. (Compare the
views of broken specimens in Figs. 17-24 to 17-26.)
3 J.
Aeronaut. Set., vol. 11, No. 2, April, 1944. See also an extensive report by
Jellinek, Latter, Thomsen, and Dorn devoted to plastic flow and fracture of
metals (P.R.R.D. report W-200, May, 1945).
214 THEORY OF FLOW AND FRACTURE OF SOLIDS

and the latter in six straight lines inscribed in this ellipse as indicated in
Fig. 15-31.
/. Mohr's Theory of Strength. The condition under which solid mate
rials begin to deform permanently or a stressed body breaks according to
Otto Mohr1 can be formulated in a general manner. A material may fail
either through plastic slip or by fracture when either the shearing stress t
in the planes of slip has increased to a certain value which in general will
depend also on the normal stress a acting across the same planes or when
the largest tensile normal stress has reached a limiting value dependent
on the properties of the material.
On the surface of test specimens of
polycrystalline ductile metals or natural
rocks which have been deformed slightly
just beyond the plastic limit fine traces
of two systems of slip lines can be de
tected on closer examination which inter
sect each other at a constant angle. On
specimens which have been strained under
a state of homogeneous stress these two
systems of isogonal lines correspond to
two parallel systems of slip planes which

...
Fio. 16-31. Constant octahedral are symmetrically inclined with respect to
shearing stress (ellipse). Constant the directions in which the major and the
maximum shearing stress (hexagon). . ,
minor principal stress <r\ and <r3 acted,
assuming that <ri > <n> <r3. The two plane systems intersect each other
along the direction in which the intermediate principal stress <r2 acted. As
a rule, the obtuse angle between these two planes of slip is bisected always
by the direction of the algebraically largest (<r1) and the acute angle by
the direction of the algebraically smallest (<r3) of the three principal
stresses.2 It is furthermore well known that specimens of brittle crystal
line materials, such as natural rocks or cast iron, or of certain brittle con
glomerates of materials (concrete) in the ordinary axial compression test
initially break along surfaces obliquely inclined with respect to the direc
tion of compression. These fracture surfaces are inclined at an angle
always smaller than 45 deg with respect to the direction of compression.
The limiting condition of failure postulated by Mohr and the orienta
tion of the two systems of the planes of slip may, in a convenient way, be

1 Abhandlungen, aus dem Gebiete der teehnischen Meehanik," 2d ed., p. 192, W.


Ernst und Sohn, Berlin, 1914; and Z. Ver. deui. Ing., vol. 44, pp. 1524-1530; pp.
1572-1577; 1900; vol. 45, p. 740, 1901.
2 See the photographs of deformed specimens
reproduced in the Figs. 20-4 and 20-5.
LIMITING STATES OF MECHANICAL STRENGTH 215

discussed by making use of the graphical representation of the states of


stress by their three principal Mohr stress circles in a plane figure (see
Chap. 10). Any point P in this plane having the rectangular coordinates
a and r defines a definite value of the normal and shearing stress a and t
in some given plane section of the stressed body. Suppose that <r and t
represent the components of stress in a plane ready to slip. Suppose that
the shearing stress t according to Mohr, in a plane of slip depends only on the
value of the normal stress a acting in the same plane, r must be a function
t = f(a) which may be represented by a curve drawn in the <r,r coordinate
plane as the locus of the points P{c,t) defining the limiting values of both
components of stress in the slip planes under the different states of stress
o\,o%,oz to which the material may be subjected. A limiting state of
stress ffi,<T2,<rj may be represented in the <t,t
coordinate plane by its three principal stress
circles. Since the intermediate principal stress
at according to the assumption made by Mohr
is without influence on the failure of a material,
it will be sufficient to consider only the major
principal stress circle corresponding to c\ and
<r3- This circle must pass through a chosen
point P{<t,t) of curve r = /(<r) defining a limit
ing state. An infinite number of such circles
IO' " " enve ope"
may be constructed through point P. The one
which is tangent to the limiting curve t = f(o) in point P is the one
representing the state of stress causing slip at the prescribed values of a
and t; for if this major stress circle were not tangent to the limiting curve,
it would intersect it and there would exist along it values of <r,r or stresses
in other plane sections of the stressed body for which the shearing stress
t at some given value of a would be larger than its experimentally observ
able value sufficient to produce slip. The limiting curve t = f(a) must
therefore be the envelope of all largest principal stress circles and the points
of tangency of the major stress circles along the enveloping curve t = f(o)
must represent in their abscissas a and ordinates r the normal and the shearing
stresses in the planes of slip. Thus, according to Mohr, the largest prin
cipal circles representing the states of stress at the limits of plasticity or
at the limits of fracture have an enveloping curve which is represented in
Fig. 15-32 by the two heavy lines.
An analytical expression for this condition is if <n > <n > a, that

(<ri c3)/2 at the limit must be a function of (<n + ff3)/2 or that the
radius of the largest principal stress circle (which is the corresponding
value of the maximum shearing stress) is a function of the distance of the
center of this circle (not shown in Fig. 15-32) from the origin 0:
216 THEORY OF FLOW AND FRACTURE OF SOLIDS

SLp j

+
=

(*L
.

fi)
(15-22)

Suppose that we introduce the equation of a major principal stress


circle in the form derived in Chap. 10:

+ = Wij^A*

r2
(15-22a)

and that we denote by

= T- - = • (15"226)
v

.
Equation (15-22a) or
(<r - VY t* = TJ

+
(15-22c)

represents the family of major principal stress circles in parameter form.


The equation of their envelope obtained by partially differentiating
is
Eq. (15-22c) with respect to the parameter p:
* = + rmTm' (15-22d)
V

where t„' = drm/dp =


(IF/dp and by solving Eq. (15-22c) for r after
substituting Eq. (15-22d) in (15-22c):
= rm VI - (r»T (15-22f)
r

Equations and (15-22e) express the equation of the Mohr enve


(15-22rf)
lope of the major principal stress circles or of the limiting curve = f(a)

t
in parameter form.
Or we may consider &3 as an independent parameter and <r\ = <p(<r%)
according to Eq. (15-22) as a function of <n.1 Now, after partially differ
entiating Eq. (15-22a) with respect to <r3 and after writing for

— d<ri — d<p
,
1

d<r3 da3

we obtain similarly as before the second, equivalent form of the equation


of the limiting curve

a = =
±

(15-22/)
t
,

Equations (15-22c/) to (22/) may serve conveniently for expressing


in

analytic form the equation of the limiting curve or of the Mohr envelope
in special applications.

See paper by C. Torre in Oesterr. Ing. Arch., vol. Nos. p. 36, 1946; also
1

2,
1,
I,
a

footnotes on pp. 407 and 408.


LIMITING STATES OF MECHANICAL STRENGTH 217

We have seen that, in general, under any plastic state of stress the
obtuse angle between the two planes of slip is bisected by the direction
of the algebraically largest (at) and the acute angle (<t>) by the direction

NORMAL TO SLIP
PLANE II

Fig. 15-33. Fio. 15-34.


Fios. 16-33 and 15-34. Planes of slip 11 and 22 and Mohr envelope.

of the algebraically of the three principal stresses.


smallest (a3) In the
Mohr representation of a state of plane stress having the principal stresses
<t\, 03 the angle 4> at which a radius

CP of a point P of the principal


stress circle is inclined with respect
to the abscissa axis a (Fig. 15-33)
is equal to twice the angle under
which the normal to the corre
sponding plane section in the
material is inclined with respect to
the direction of the principal stress
<ri (Fig. 15-34). We see that the
<tt>

angle of inclination ^OCP^ =


of the normal P\C to the envelop
ing curve =
/(<r) in the point of
t

tangency Pi with major stress


a

circle in Fig. 15-33 must be equal


<t>

to the acute angle between the


two planes of slip corresponding
to the state of stress <r\ > a* > <r3. Fio. 15-35. Slip line pattern on face of a
paraffin prism after test in axial compres
Mohr's theory explains and ac sion parallel to long side of prism.
counts for an interesting fact,
<tt>

namely, that the angle between the two surfaces of slip for a given plastic
state of stress a\,ai,<r3 in general depends on <r1 and a3, that may change
it

from material to material and also under work hardening. In ductile


218 THEORY OF FLOW AND FRACTURE OF SOLIDS

metals (mild steel) the angles under which the slip planes (thin plasticized
layers of material) are inclined with respect to the direction of the major or
minor principal stress tend to approach in some brittle materials
45 deg while

(such as marble or sandstone) the angles of the two slip planes with the
principal direction of o-3 may differ at fracture considerably from 45 deg

Fig. 15-36. Flow lines on a circular plate of steel deformed by a load in center. (After
L. Hartmann.)

(<t> < 45 deg).Furthermore, in the former cases the tensile and compres
sive strength at the yield point become equal (mild steel) and Mohr's
"enveloping curve" degenerates into two straight lines parallel to the
a axis. Mohr's theory coincides in this case with the theory of maximum
shearing stress (Sec. 15-lSe).1 In the second case, on the contrary, the
1 Mohr in one of his original papers drew the two envelopes of the major stress
circles in such a manner that they would intersect the positive a axis symmetrically
in an oblique direction. A. Leon (cf. p. 221) pointed out that this would lead to
contradictions concerning the limiting conditions represented by points of the enve
lope in the region of tensile stresses and the experiments on tensile fracture. When
the envelope intersects the a axis in a perpendicular direction, however, these contra
dictions are eliminated.
In summing up, we may state that the Mohr theory predicts many of the principal
facts known from strength tests for a large variety of materials embracing the per
fectly ductile metals and the solids behaving in a brittle manner, including the obser
vations on the orientation of the surfaces of slip in plastic materials and, as we shall
see below, of the shear and cleavage fractures. In the region of large mean pressures
LIMITING STATES OF MECHANICAL STRENGTH 219

theory may express the behavior of such materials for which the yield or
fracture stress in tension differs considerably from that for uniaxial
compression.
A special case of the Mohr theory of strength of considerable practical
interest in various applications is obtained when the enveloping curve
consists of two straight lines inclined at the same angle with respect to the
<r axis; Eq. (15-22) then takes the form


o~\ &3 cl + c2(<ri + a) . (15-23)

This limiting condition of failure is expressed also by stating that a


material does not fail unless the
shearing stress t in the plane of
slip becomes equal to the right side
of the inequality:

±t, + na (15-23a)

where a denotes the normal stress


acting across the plane of slip and
ti and n are given positive con
stants. For both upper or both
lower signs and for the equality
sign, these are indeed the equa
tions of two straight lines connect
ing the points P having the coordi
nates a and t in which the major
stress circles are tangent to the two Km;. 15-37. Charles AuKU.it in Coulomb
straight lines which were postulated (1736-18,6). (From " Great Men of Sci
for the envelopes. Equation ence," by P. Lcnard,
pany, New York, 1933.)
The Macmillan Com

(15-23o) was proposed by Coulomb


for expressing the condition of rupture in simple compression of a brittle
material. According to C. Duguet,1 a brittle material breaks in biaxial
compression (under moderately large values of the two compression
stresses) in a shear fracture when the shearing stress in the plane of
fracture reaches a critical value t which depends on a certain shearing
stress ti but increases proportionally with the value of the normal stress a
if this is a compression stress < 0). The term na in Eq. (15-23a)
(<r

when the envelope tends to become parallel to the a axis, predicts what might be
it

called "pure" shear fracture (</t>= 45 deg), in the transition zone the ordinary obliquely
<t>

inclined shear fracture (0 < < 45 deg), and near the apex of the envelope the cleavage
fracture (</>= 0) in the direction normal to the maximum tensile stress.
Duguet, C, "Limite d'6lasticit6 et resistance a la rupture," partie, 1885.
1

2
220 THEORY OF FLOW AND FRACTURE OF SOLIDS

is similar to a friction force, and the proportionality factor fi may be


interpreted as a coefficient of internal solid (Coulomb type) friction.
Equation (15-23) was proposed by J. J. Guest1 for the condition under
which the ductile metals start to yield. From his tests published in
1900, he was led to assume that the first term in Eq. (15-23) containing
the material constant Ci (corresponding to the maximum shearing stress
theory) does not express the condition of flow with sufficient accuracy
and that a second term is required for obtaining a better agreement
between Eq. (15-23) and the test results. According to Guest, the con
stant c2 in Eq. (15-23) has a comparatively small value so that the slope
of the two straight envelopes is small for expressing the condition of yield
ing for ductile metals.2
Instead of using the representation in a plane <t,t one may define the
"surface of yielding" in the three-dimensional way described in the
previous sections. The surface of yielding in a ai,<rs,o-» system of coordi
nates is represented by a slender prismatic body having regular hexagons
for its sections normal to its axis but with variable lengths of their edges
for the Mohr theory and by a slender hexagonal pyramid for the Guest
theory, both these surfaces being oriented so that their axes are inclined
at equal angles with respect to the cri,<T2,a3 axes.3
The envelope of the major principal stress circles representing limits
of yielding must satisfy further requirements. Clearly, it cannot inter
sect the a axis on its negative side, since this would amount to assuming
that solid materials can be made to deform plastically under pure hydro
static pressure (en = <rt = <r3 = — p, p > 0). Experimental evidence
indicates that the envelope for large negative values of er tends to bend
in a direction parallel to the a axis. On the other hand, if the two sym
metric branches of the Mohr envelope intersect at a point of the positive
a axis by crossing it at a finite angle, this means that the major principal
stress circle tangent to both branches of the envelope at this point has
contracted to a mathematical point. Again, if the envelope represents
the condition of yielding, this contradicts the experimental facts, because
the triaxial state of three equal tensile stresses, under which a material
cannot yield, would correspond to such a point. This led investigators
to assume that the Mohr envelope for yielding loses its physical meaning
as it approaches the positive a axis.
1 Yield Surface in Combined Stress, Phil.
Mag., vol. 150, p. 261, 1940. Also Phil.
Mag., vol. 50, p. 69, 1900 (containing the original test results on which Guest bases
his equation).
1 In terminology used by Guest, this amounts to introducing the "volumetric"
term Cj(<r, + a,) in Eq. (15-23).
'See footnote, p. 208; Mehldahl illustrated the Mohr theory among his models of
these various surfaces.
LIMITING STATES OF MECHANICAL STRENGTH 221

Since the limiting surface of yielding f\(<r\,(m,<ri) = 0 (Sec. 15-1) cannot


intersect the straight line along which ax = o-2 = <r$ in a rectangular
system of coordinates <ri,<rj,<r3 for ductile materials in the positive quad
rant of <T\,<ri,a3, a second surface distinct from the first one must exist,
namely, the surface of rupture fi(<ri,<ri,ag) = 0 (Sec. 15-3) which contains
the points 0-1,0-2,73 representing the limiting states of stress under which
the tensile-type (cleavage) fractures occur. Investigators have assumed
that the surface of yielding /i(<ri,<ri,os) = 0 cannot be extended in the
positive quadrant beyond its intersection curve with the surface of frac
ture ft(ai,(T2,<rs) = 0. The opinion was also expressed that the fracture
of brittle materials by tension or under states of stress with two or all
three principal stresses positive does not follow the laws which, according
to Mohr, postulate the existence of an envelope of the major principal
stress circles.
Alfons Leon in Vienna1 further clarified important additional conditions
as far as the fracture of brittle mate
rials is concerned by pointing out
that a Mohr envelope has a physical
meaning in the region near which it
may intersect the positive a axis in aa,r
plot if it intersects it at.a right angle
and if the enveloping curve of the
major principal stress circles has a
Fio. 15-38. Fig. 15-39.
finite radius of curvature at this point. Figs. 15-38 and 15-39. Mohr envelopes.
Suppose that a major stress circle
just coincides with the circle of curvature of the Mohr envelope at its
apex A (Fig. 15-38). Since the normal to the envelope at a point P(<t,t)
for P = A,
<j>

r = 0. This corresponds
<£,

defines the angle of slip =


0,

to a cleavage fracture, whose direction normal to the major principal


is

stress a\. All major stress circles tangent to the apex whose diameters
A

are smaller than the diameter of the circle of curvature in will represent
A

possible states of stress under which cleavage fractures must occur normal
to c\. Depending upon whether the diameter AB of the circle of curvature
at the apex of the envelope smaller (Fig. 15-38), equal to, or larger
A

is

(Fig. 15-39) than the distance OA, being the origin of the <t,t system,
0

the material under simple tension (<n = o-3 = will break in an


0)
0,
>

tr2

Ueber die Rolle des Trennungsbruches im Rahmen der Mohr'schen Anstren-


1

gungshypothese, Der Bauingenieur, vol. 15, No. 31, 32, 1934. Also Proc. 4th Intern.
Congr. Applied Mechanics, Cambridge, England, 1934; Ing. Arch., vol.
4,

p. 421,
1933. For an instructive survey of the limiting conditions of flow and of fracture the
reader also referred to paper by L. Rendulic, Eine Betrachtung zur Frage der
is

plastischen Grenzzustaende (Der Bauingenieur, vol. 19, pp. 913-916, 1938).


222 THEORY OF FLOW AND FRACTURE OF SOLIDS

oblique fracture (when AB < OA) or by cleavage (AB ? OA), since the
major stress circles for simple tension in the former case touches the
envelope at a point Pi, for which t > 0, <p > 0. Consequently, accord
ing to Leon, if the Mohr enveloping curve t = f(a) has such a form that
AB < OA only those states of bi- or triaxial tensile stresses will produce
cleavage fractures whose major stress circles are smaller than the circle
of curvature in the apex A of the envelope. It is interesting to note that

is,
the sharper the curvature of the envelope in its apex the more the
fracture law will deviate from the Rankine maximum stress theory in the
direction postulated by McAdam's fracture tests which were quoted on
page 181. However, in this case state of simple tension

a
ffl > = =

0,
(Tj Cz

0
should produce an oblique (shear) fracture and this plane, as point Pi
approaches A, should turn into the direction normal to the major stress
v\. It may be that
some of these facts may prove helpful also for explain
ing the cup and cone fracture of ductile metals after considerable cold
deformation. Thus we see that a Mohr envelope having a finite radius
of curvature at its apex in general includes states of stress under which
cleavage as well as the shear fractures can occur in a material.1

If the circle of curvature at the apex of the envelope large enough, may-

it
is
1

include the major stress circle corresponding to a state of pure shear <ri = — a», *t = 0;
this would represent brittle material for which a cleavage fracture should be observed
a

both in a tensile and in a torsion test with a round bar which the case for marble or
is

cast iron. On the contrary, the circle of curvature in just passes through the
A
if

origin but the major circle for pure shear touches the envelope outside of point A,
O

this would define a material which breaks in simple tension by cleavage normally to
the tension stress but in torsion by a shear fracture.
Leon has discussed the properties of an envelope for which he assumed an ordinary-
parabola t' = Ci This enabled him to fit the constants to suit brittle materials
+

csff-
for which the ratio of the compression to the tensile strength reaches large values of
the order of or more, such as cast iron, concrete. See his papers on cast-iron
5

strength ("Mitteilungen des technischen Versuchsamtes," Vol. 22, p. 17, Vienna,


1933) and on concrete ("Beton und Eisen," Vol. 34, No.
8,

1935).
The appearance of the rupture surface in polycrystalline brittle materials in a
"shear" fracture has frequently been contrasted with that of "cleavage" fracture
a

by observers. The former one appears usually covered by debris of detached par
ticles and shows signs of having been polished in certain portions while the latter
is

described as perfectly smooth. It should be noted, however, after considering the


details which were just discussed, that this perhaps superficial observation. The
is

conditions of fracture surface which inclined at a finite angle with respect to the
is
a

principal direction of <ri and <r3 depends on one important fact (attention to which
has not been given), namely, on whether the normal stress <r in or
is
it

compression
a

a tensile stress. The majority of "shear" fractures described in the literature refer
to those cases for which a < 0. It interesting to note that, according to the
is
LIMITING STATES OF MECHANICAL STRENGTH 223

A study of the Mohr envelope in the region where it intersects perpen


dicularly the a axis in the <r,r rectangular plot of the limiting states of
stress as emphasized by Leon may throw some very interesting light on
another of its useful possible applications to which special reference
should be made here, namely, in geology. Geologists have been carefully
studying the "faults" and the "jointing" in rocks.1 They noticed long
ago that faults of certain types may be described as a double system of
parallel planes (shear faults) in which the rock strata on one side of the
fault have been displaced relative to those on the other side; furthermore,
that in certain very solid rocks double and even multiple systems of
parallel cracks (joints) may be seen traversing the rocks. The phenom
ena of shear faults and of the jointing seem to be manifestations of the
formation of the surfaces of slip or of cleavage cracks on a grand scale in
nature. Since in their study comparatively little use has apparently
been made of the facts mentioned above which the Mohr theory disclosed
in stressed solid materials, attention is drawn to the possibility of pre
dicting and describing the mechanical laws of the formation of faults and
joints in rocks by basing them on this or related concepts developed in
mechanics.
It that no satisfactory mechanical explanation has yet been
seems
advanced, for example, for the formation of joints in rocks (of igneous
origin, in granite, in which no stratification is observable) appearing
frequently with a remarkable regularity as a dense system of parallel
cracks or of planes of weakened strength along which the rocks may easily
be separated or appear definitely broken. The distances between these
cracks may be measured in inches or fractions thereof or feet or yards or
by larger intervals, etc. Remembering the facts brought out in the pre
ceding remarks, the preceding theory may predict, assuming certain special
curves as Mohr envelopes, a double system of parallel surfaces of weakness,

thoughts developed above in the text, oblique fractures may quite gradually develop
<tt>

<t>

under increasing angles beginning from the case = of what was characterized
0

(according to the above distinction) as a pure cleavage fracture. These oblique


<t>

fractures inclined at finite angles cannot show the damaged structure containing
debris of the grains of the shear fractures the normal stress in them was positive and
if

a tensile stress. Thus a gradual transition a pure cleavage to a shear fracture in con
of

trast to the necessity sharply distinguishing between two conditions such asf(<ri,<Tl,a3) =
of

and /2(<ri,<r2,<r3) = must be conceded.


0

B., and R. Willis, "Geologic Structures," 3d ed., p. 121, McGraw-Hill


Willis,
1

Book Company, Inc., New York, 1934, 544 pp. Longwell, C. R., A. Knopf, and
R. F. Flint, "Geology," 2d ed., Part Physical Geology, p. 332, John Wiley Sons,
&
I,

Inc., New York 1939, 543 pp. Cloos, H., "Einfuehrung in die tektonische Behand-
lung magmatischer Erscheinungen Granit Tektonik," Verlagsbuchhandlung Gebrtlder
:

Borntraeger, Berlin, 1925.


224 THEORY OF FLOW AND FRACTURE OF SOLIDS

symmetrically oriented to the directions of principal stress in which fractures


disclose very nearly the characteristics of pure cleavage cracks; furthermore
this theory predicts angles at which both parallel systems of surfaces are
inclined relative to each other which may vary between zero (when both systems
coalesce in a pure cleavage crack) and a right angle (when they are the slip
planes of a perfectly plastic substance). Pure tensile stresses in rocks
buried underground may seldom be encountered in nature if at all. The
element of time, however, may be a most important factor influencing the
gradual spreading of joints and the splitting mechanism in the jointing of
rocks over geological periods of time.
Experimental evidence which has become available during the last two
decades on the condition of yielding of ductile metals indicates that one
assumption of the Mohr theory is not verified by the experiments, namely,
that the intermediate principal stress should have no influence on the
condition of yielding [see Eq. (15-22)]. This has led various investigators
to propose a broader formulation of the limiting condition of yielding
established by Mohr's Eq. (15-22) by also including the value of the inter
mediate principal stress <ri on the right side of the flow condition [Eq.
(15-22)].
According to Mohr's assumptions, the shape of the characteristic
enveloping curve of the principal stress circles for all limiting states of
stress producing yielding should not depend on the intermediate principal
stress. If, for example, a material in tension under the same
yielded
stress as in compression, so that the largest principal stress circle for both
cases (pure tension, pure compression) had the same diameter, the yield
stress in pure shear should be equal to one-half the value of the yield stress
either in tension or in compression. This has not been verified, since
recent tests made with such materials show a ratio considerably higher
than one-half.
g. Conditionof Slip in a Loose Granular Material. If the constant ci
in Eq. (15-23) is taken equal to zero, this equation expresses a limiting
condition, which coincides with the limiting state of equilibrium encount
ered in the plane problem of the distribution of the stresses in a hose granu
lar material such as sand. Such a material cannot transmit tensile
stresses; it has no cohesion. The two straight lines representing Mohr's
enveloping curve must therefore intersect each other at the origin O of
the plane, and they have a physical meaning only for compression
<t,t

stresses. Mohr made use of this representation of the limiting states of


equilibrium in a theory of earth pressures on retaining walls. The equi
librium of granular material is characterized by the condition that
a loose
the shearing stress in any plane section can never become larger than the
friction force caused by the normal pressure in this section when slip
LIMITING STATES OF MECHANICAL STRENGTH 225

starts in the mass. Assuming Coulomb's law of friction according to


which the friction force is proportional to the normal force, the constant
Ct in Eq. (15-23) corresponding to this case may easily be expressed in
terms of the friction angle p or of the coefficient of sliding friction between
the grains p, = tan p.
If ffi and 0-3 denote the major and minor compressive principal stresses
(taken positive), the limiting condition
of slip [Eq. (15-23)] is expressed by

o\ — oz = sin p ((Ti + <rs) (15-24)

since in Fig. 15-40 CP = (<n


- <r,)/2,
OC = (<7i + <r,)/2, and
sin p - CP/OC .
Fig. 15-40. Mohr envelope for
The points of tangency P of a major limiting equilibrium of a loose
material.
stress circle and the straight line UA
representing one of the two envelopes of the major stress circles have the
ordinate PQ = t and the abscissa OQ = a and the limiting condition of
Coulomb friction r/<r = p. = tan p is indeed satisfied if the angle AOC is
taken equal to the friction angle p. Obviously all points situated within
the sector AOB have ordinates (equal to the shearing stress t) which in
their absolute values are smaller than the limiting shearing stresses at
which slip becomes possible, namely, t = pa.
h. The Octahedral Shearing Stress r„t Is a Function of the Mean Normal
Stress aw. We saw in the preceding section that, according to the Mohr
theory of strength, the shearing stress t at the plastic limit in the planes of
slip was dependent on the normal stress a or that t = /(<r) defined the
condition of flow and an enveloping curve of the major principal stress
circles. This condition was based on the hypothesis that the intermedi
ate principal stress o-2 would have no influence on this condition since the
normal stress <n acts in a direction which is parallel to the planes of slip.
The condition of yielding may be formulated more generally by referring
it to the octahedral planes of the stress tensor instead of to the two planes
of slip. This amounts to assuming more generally that the octahedral shearing
stress Toct defined by Eq. (15-17) at the limit of yielding is a function of the
octahedral normal stress a„i. Since o-oot is equal to the mean stress
(0-1 + a + <7s)/3 the equation


Toot /(foot) (15-25)

expresses a condition of flow in solids in which the limiting value of the


shearing stress Toot in the octahedral planes depends on the mean normal
226 THEORY OF FLOW AND FRACTURE OF SOLIDS

stress (<n + <n + (T3)/3 and thus also on the value of the intermediate
stress <T2.

This equation represents a surface of revolution in a system of rec


tangular coordinates, whose axis is
equally inclined with respect to the co
ordinate axes <7i,<rj,ff3; for, if we refer the
equation of the surface of yielding
fi(<r 1,0-2,0-3)
= 0 given by Eq. (15-25) to a
new system of rectangular axes <ri',<ri,<f*
in which, for example, the positive <73'
axis coincides with the space diagonal
%■ ffi = 02 = ff3 of the first system and if we
chose the and oV axes in the plane
<r\

01 + <n + o3 = 0 (which is normal to the


straight line 0-1=0-2 = 03 and passes
Fiu. 15-41. vi, o-!. az through the origin 0) in two convenient
and oY, a,',
a,' axes. {<ri'\[CD, <rj\\AB, o-i'j.
perpendicular directions (see Fig. 15-41),
the equations for a transformation of the
coordinates may be expressed as follows:

01 , ffs . 03

" Pi ff2 1 O3
'
(15-26)
V6 -\/2 V3
0-.3

and the expressions for the components of stress Toot and 0«„t are trans
formed into

- - - -
(<r,"

1 1
.* [(*, + o,)2 + = + <r,'A
<r2)2 (02 (03 <n)']

(15-27)
= - + =
+

(o-l <T2 O3) —=■


V3
3

Equation (15-25) may be also written in the form

o,'2

--'(3)
+

02 = f>/2
+

or <n' <T2/2) (15-28)

representing in the system oV, 0-2', 03' the equation of a surface of revolution.
If, for example, a circular cone assumed as the surface yielding, Eq.
is

of

expressed explicitly by
is

(15-25)
LIMITING STATES OF MECHANICAL STRENGTH 227

9w = 2(3coao.. - c,)2 (15-29)


or by

3(<n'2 + a*'*-)
= 2(V3 c^' - c,)2 (15-30)
where

Co
= •— -r— ' ci = —— (15-31)
ae + at ae +i ai

are two material constants in whose expressions ac denotes the yield stress
in simple compression and at in simple tension. The apex of the cone
has the distance

2acal
(15-32)
V3 (a*

at)

from the origin 0 of the ahat,at axes and the tangent of the angle a of the
generatrices of the cone with its axis is given by

tan a =
V^-*)ac ■+■at
(15-33)

If the surface of yielding is a paraboloid of revolution, Eq. (15-28) becomes

at' = a, + a2(<n'2 + at'2) (15-34)

Equations (15-30) and (15-34), respectively, may represent con


veniently solid materials for which the yield stresses in simple tension and
compression are different.1 On the other hand, they may represent the

1 It is noted again that in all cases when the surface of yielding intersects the space
diagonal ax = at = a3 a certain portion of this surface must be excluded around the
points of intersection since a material cannot yield under a triaxial state of three equal
tensile stresses. A condition of flow related to Eq. (15-25) has been suggested by
F. Schleicher (Z. angew. Math. Afechanik, vol. 6, p. 216, 11)26), who assumed that
the total elastic strain energy at the limit of plasticity is a given function of the mean
stress a = (en + at +<rj)/3.
The total elastic strain energy or the elastic energy of distortion (see Sec. 15-13d)
stored in a unit of volume can be expressed by quadratic functions containing the
first or the second invariants of the stress tensor (see p. 92). These criteria of failure
based on energy relations thus postulate certain functions connecting these two
invariants. Since no strict physical reasons can be advanced why only these two
particular "stress invariants" should be significant, we prefer to formulate the condi
tions of failure on a phenomenological ground, as quoted in the text above.
We may note that the two surfaces of revolution [Eqs. (15-30) and (15-34)] when
expressed in terms of the three principal stresses ai,a2,a3 have equations of the form
Cone:
(<n
- <r»)' + (<n
- <r.)« + (<r3
- <n)»
=
W[co(ffi + a, + a,) - r,]« (16-30o)
Paraboloid:
ax + a, + a, = 3a, + 9ai[(«ri
- <r2)2 + (a. - <r3)5 + (<r3
- a,)'] . (15-34a)
228 THEORY OF FLOW AND FRACTURE OF SOLIDS

surface of rupture for brittle materials for which the compression strength
differs considerably from the tensile strength.
If the surface becomes a circular cylinder and we obtain
ac = at = <to
from Eq. (15-20) again Eq. (16-17) representing the condition t<*a = const,
under which an ideally plastic material flows.

They may represent generalized plastic substances for which the yield stress (measured
by the intensity of the octahedral shear stress Toct) is a function of the mean stress or
pressure ow = (<ri + <r2 + o-3)/3. One or the other of the last two equations may
be set aside for developing the theory of the corresponding substance in manner
analogous to that of the ideally plastic substance (for which Toot = const; see Chap.
27) and the general theory of the surfaces of slip, e.g., in the two-dimensional cases of

flow, may be investigated based on the characteristic curves of the corresponding


systems of differential equations (see Chap. 37).
Equation (15-34a) and some of its applications were discussed in a noteworthy
paper by C. Torre (Die Grenz-Zustande statisch beanspruchtcr Stoffe, Schweiz.
Arch, angew. Wiss. u. Tech., vol. 15, Nos. 4, 5, 1949).

S
CHAPTER 16

CONDITIONS OF FAILURE. RULES IN THE NEW THEORIES


OF FLOW OF SOLIDS

Before reporting the results of recent experimental investigations on the


limits of plasticity in various materials, on the laws of yielding of ductile
metals, and on fracture it seems in order to clarify a few concepts on which
to base comparisons of the test results. During the second half of the
nineteenth century investigators concentrated their efforts on a formula
tion of the condition under which solid bodies reach the plastic limit or
break, i.e., on defining only the condition for the limits of plasticity or of
fracture. Since the second decade of the twentieth century, considerable
attention has been concentrated beyond the condition of plasticity on the
mathematical formulation of a complete set of equations for describing
the steady slow flow of plastic masses or the permanent distortion of solid

bodies.
When summarizing the ideas leading to the various theories of strength
which were listed in the preceding chapter, we may perhaps state a few
conclusions:
1. The constant maximum shearing stress and the constant octahedral
shearing stress theories seem to express in a satisfactory manner the con
ditions under which the ductile metals start to yield permanently.
According to both theories, the stress under which the material starts to
yield in simple tension or compression should be equal.
2. Under a state of pure shear (<ri = r, <n = 0, a3 = — t) there is a
marked difference between the limiting values of r predicted by both
theories. When Tma« = co/2 = const, at the limit of plasticity in a state
of pure or simple shear (torsion test) t = <to/2 while when

T«t = ——
V2<r0 =
o const ,

t = ffo/\/3 = 0.577o-0, <ro being the yield stress in simple tension. For a
state of pure shear the intermediate stress <r2
= (<ri + <r3)/2 = 0. Gen
eral states of stress a\ > <n > a3 in which a* = (o\
+ <r3)/2, i.e., for which
the two minor principal Mohr circles have the same radii, will be par
ticularly suitable for determining experimentally which of the conditions
Tia„ = const or Toct = const is verified by tests or whether the intermediate
229
230 THEORY OF FLOW AND FRACTURE OF SOLIDS

principal stress <r2 has no influence or has an influence on the condition of


yielding and on the radius of the largest principal stress circle (v\ — <r»)/2-
3. When a ductile metal breaks in a shear fracture after having been
deformed permanently under biaxial states of stress, the maximum shear
ing stress Tmmi = const may express the condition of such a fracture.
4. The behavior of the natural rocks which under tensile stresses
behave like brittle materials when they are brought in the plastic state
under combined compression stresses, i.e., under the conditions of loading
which may produce plastic deformation in them is satisfactorily predicted
by the Mohr theory.
5. The cleavage and shear fractures in brittle polycrystalline materials
(rocks) under a moderately large mean stress (<ri + <r2 + *«)/3 seem to
follow the conditions specified by the Mohr theory.
6. The assumption of an envelope of the largest principal stress circles
according to Mohr (or the maximum shearing stress or the Guest theories)
from a mathematical viewpoint, however, involves considerable difficulty
with respect to a formulation of the equations of flow in ductile materials
under general states of stress. R. von Mises1 and H. Hencky2 have
avoided the mathematical difficulties by also including the intermediate
principal stress in the condition of flow. After recalling that the limiting
f
condition of yielding 1(0 1,01,03) = 0 may be represented by a surface in a
system of rectangular coordinates 01,02,03, we may add that the theories of
flow based on conditions of slip, the Mohr, the maximum shearing stress,
or the Guest theories, in which the hypothesis is made that the value of
the intermediate principal stress <r2 (when 0\ > 01 > 03) has no influence
on the condition of yielding, lead to six distinct conditions of flow (in
which, however, always one of the three principal stresses does not
appear) and are represented consequently in a 01,01,03 system of coordi
nates by surfaces having six sides. The discontinuity inherent in the
form in which the conditions of slip and of yielding were assumed hy
Mohr leading to surfaces of yielding with six edges was eliminated by von
Mises when he proposed replacing the regular hexagonal prism represent
ing the maximum shearing stress theory by the surface of the circum
scribed straight circular cylinder expressing also the condition of a con
stant octahedral shearing stress (Sees. 15-13d and 15-13e):

1 Mcchanik dor fosten Korper im plastisch-deformablen Zustand, Nachr. Gc3. Wist.


Gottingen, Math.-physik. Klasse, 1913.
1 Zur Theorie plastischer
Deformationcn, Z. angew. Math. Mechanik, vol., 4, p. 323,
1924; and Lvber das Wcsen der plsistisehen Verformung, Z. Ver. deut. Ing., vol. 69,
p. 695, 1925. In the discussion of this condition of plasticity at the First International
Congress for Applied Mechanics in Delft (1924) it developed that M. I. Hubcr (Ix>m-
berg) had also independently suggested Eq. (16-1) as the condition of yielding.
RULES IN THEORIES OF FLOW 231

Too.' =
\ [(<ri
- <ra)2 + (<r2
- <r3)2 + (<r3
- <r,)2] =
^f
= const . (16-1)

7. Equation (16-1) assumes that the limiting value of the shearing


stress T„ot does not depend on the value of the mean stress a or pressure
p = —<T = — (<ri + <ri + <r3)/3 under which the plastic distortion starts.
This is particularly the case in the ductile metals. However, in other
solid materials (rocks) cases are known (Chap. 18) in which the condition
of flow must preferably be assumed according to Eq. (15-25) (Sec. 15-13A)
or as follows:
=
Toot /(*) . (16-la)

In other words, in certain polycrystalline solids, unlike the ductile metals,


the shearing stress r„t under which they can deform plastically increases
with the mean pressure p.
Let us now review a few physical facts characterizing the way solid
materials deform permanently. Consider the slow continuous distortion
of a plastic mass in which the change in shape of the mass is relatively
small. In polycrystalline solids the permanent part of the distortion
may be regarded as sufficiently large so that the elastic part is negligible
compared with it.1 We also assume that at each position in the slowly
flowing mass both the value and the direction of the principal stresses
are known and that the directions of the principal stresses at a given
point do not change during the flow. In order to define the deformation
of the masses we may consider a small cube at an arbitrary point in the
body in its initial condition. We shall take the sides of this cube in the
planes of the principal stresses. The question as to what deformation
of the plastic mass under the given stress distribution is to be expected
may be formulated as follows: In which way will a small cube, the edges
of which are parallel to the instantaneous directions of principal stress in
the deformed state of the body, distort? Does the material yield in the
direction of the intermediate principal stress, and if so, what increase in
extension occurs along the corresponding edges of the small cube? The
answers to these questions may be stated in the form of three rules as
follows:
1. The directions of the small increments of the principal permanent

extensions coincide with those of the principal stresses at all times.


2. The density or the volume of the mass does not change appreciably.

1 This is the consequence of the fact that the moduli of elasticity of most nonorganic

crystalline or amorphous solids are comparatively large quantities of the order of mil
lions of pounds per square inch. In certain organic materials (the so-called "plas
tics") having long chain molecules, this may no longer be true.
232 THEORY OF FLOW AND FRACTURE OF SOLIDS

3. The rates of shear are proportional to the shearing stresses. The


figure representing the tensor of the rates of strain by its principal Mohr
circles remains continuously geometrically similar to that of the group
of the principal stress circles (Figs. 16-1 and 16-2).
Suppose that the permanent strains are comparatively small. Accord
ing to Sec. 11-2 we may then represent a state of small strain (e,7) as well
as a state of stress (<r,r) by their three principal strain and stress circles,
respectively, e is the unit elongation corresponding to some given

Fig. 16-1. Fio. 16-2.


Fio. 16-1. Tensor of rates of strain.
Fio. 16-2. Stress tensor.

y the unit shear of the planes perpendicular to this direction,


direction,
and t the shearing stress in these planes.
a is the normal stress,
We shall now consider these statements in order. The first rule is
only another expression of the experimental fact that the directions of
the largest increments of the shearing displacements coincide with the
instantaneous directions
of the maximum shear stresses in a deforming
body. We must conceive of the plastic state of solid materials as being
such that parts of the mass slide with respect to one another infini-
tesimally simultaneously along countless slip planes.
The second rule records the usual behavior of most solid materials.
To be sure, although very exact measurements of the distortion during
large plastic deformations have shown a small change in the volume (since
the microstructure with increasing distortion gradually breaks up in the
crystallites, the volume increases a little), these changes are, in the case
of metals plastically deformed, of the order of the elastic deformations
and may therefore be neglected if the permanent distortions predominate.
The second rule is expressed by the condition

t\ + «2 + «s
= 0 (16-2)
RULES W THEORIES OF FLOW 233

in which ti,t2,«3 are the principal strains, as assumed, the permanent

if,
change in shape of the mass relatively small.1

is
The third rule expresses the relative changes of the lengths of the
edges of the small cube. We know the way in which plastic mass

a
deforms under a homogeneous state of simple tension, under pure shear,
and under simple compression. Suppose that we observe an infinitesimal
distortion which measured by the increments of the principal conven
is

tional (small) strains dt\,dti,dt3. We know then that


Under simple tension the distortion given by

is
<Xl — 0°
= —

0
<7"2 0"3

,
,
dtt = dt* = — dt i/2 .

Under simple or pure shear the distortion given by

is
= = = —
0

t
<Tl <t2 0-3
t
,

,
dti = —
d(3 de2 =

0
.
,

Under simple compression the distortion given by


is

= = = — <r
0

&1 <Ti <T3


,
,

rfei = dti = —
dt3/2 .

Suppose that we divide the increments of strain by dt, by the time


required for their formation. The quantities dti/dt, dti/dt, de3/dt measure
the time rates with which the principal strains changed. These will be
very small quantities, since we assume not only small changes in shape
«i,«2,<3, but also very slow rate of flow. Solid bodies usually deform at
a

permanent rates of strain dti/dt, dtt/dt, dt3/dt which are quite small.
Like the tensor of a state of small strains ei,e2,e3 the tensor of the rates of
strain may be represented by its principal Mohr circles by plotting the
permanent rates of strain (dt/dt) as the abscissas and one-half of the
permanent rates of shear (}4.dy/dt) as the ordinates of points in a dt/dt,
}^dy/dt system of plane rectangular coordinates (Fig. 16-1). In the
figure of the principal strain rate circles the origin of the axes must be
0

chosen so that Eq. (16-2) expressing the condition of incompressibility


of the material satisfied, or that
is

dti det dt3 = „


0- nr o
.

+
lF
\

(16-2a)
■dt+W
Figure 16-1 shows a simple geometric construction by means of which the
location of the origin may be determined. We note that the three
0

the strains are of finite magnitude, the condition of incompressibility


If
1

is

expressed
by +«,)(! +«*)(! +*,) =
1.
(1
234 THEORY OF FLOW AND FRACTURE OF SOLIDS

states of plastic distortion corresponding to simple tension, simple or

pure shear, and simple compression are represented by groups of circles


geometrically similar to those of their principal stress circles.
Suppose that this rule of the geometric similarity of the figures of the
three principal stress circles and of the three principal circles representing
the corresponding tensor of the rates of strain is true in general for all
plastic states of stress whose principal stress circles are represented by
Fig. 16-1. Let Si,Si,S3 be the points at which the principal stress circles
intersect the <r axis in Fig. 16-2. In order to define the position of the
point St on the a axis in Mohr's stress plane <r,r relative to the points S\
and S3 having the distances o\ and <r3 from 0 we introduce the ratio it
defined by

=
MS,
= at -
- [(gl + g,)/2] f
" MS,
m
(ar-^/2 (16"3)

where point M is the center of the major stress circle. Similarly, suppose
that we define a ratio

, - -^
^
+Jl/2] (16-4,

where «i = dei/dl, • • • . Since

For simple tension: n = v = — 1 ,

For simple or pure shear: n = v = 0 ,

For simple compression : ft


= v •= 1

we may postulate the third rule of flow to hold for all possible stress dis
— in the form
tributions 1 ^m= + 1

or
— Cl — (16-5)
— «J
fr2 ~H Ol C.l _ «2

fl + «2
— —
<7l C3 «i «3

If the
principal stresses <ti,<t2,<t3 do not change their values at the plastic
limit when the principal permanent strains slowly increase under the rates
of strain ei,«2,«3 under the steady slow flow we may write ei = itt, «j = *»/,
e3
= i3l provided that the strains ei,€2,e3 measure the distortion from the
unstrained condition of the body at the time t = 0, and we may substi
tute in Eq. (16-5) the strains «i,e2,«3 for the rates of strain *i,e2,«a. Thus, in
the latter case if we now define alternatively
— —
«2 «i + «2 t3
(16-6)

we still have the rule that n = v.


RULES IN THEORIES OF FLOW 235

To define three mutually perpendicular directions in space we need to


specify three angles. The first rule, according to which the principal
directions of stress <ri,<ri,<r3 must coincide with those of the principal rates
of strain ii,ii,i3, requires that three conditions be satisfied. Each of the
second and third rules adds one condition more to these. Thus, in the
three laws of yielding, five independent conditions are involved, which
must define the stress-rate-of-strain relations under which the plastic
distortion of solids develops.
The third rule of flow may be expressed in two alternative forms. We
have
The principal stresses: <r\,ai,a3
The principal shearing stresses: ti,t2,t3

Tl = (<t2

<r3)/2 ,
= —
T2 (<T3 <rl)/2 ,

Ti = (<Tl <r2)/2 .

The principal rates of strain: «i,e2,e3


The principal rates of shear: 71,72,73

71
= «2

e3 ,

72
= «3

«i ,

73
= «i — «2 .

For the principal shearing stresses the identity holds:


Tl + + = 0
T2 T3 . (16-7)

For the principal rates of shear we have the identity


+ + = 0
71 72 73 . (16-8)

Equation (16-5) may be split into

a-2^1
a1 — <T3
= i^Jl
«1

«3
and °JL^!=i±zJl.
a1— —
<T3 tl ti
(i6.9)

After substituting the variables ti,t2,t3 and 71,72,73, assuming that the
two identities Eqs. (16-7) and (16-8) hold, we see that rule 3 can be
expressed in a second form:

71:72-73 = ti:t2:t3 . (16-10)*


* Thpse are not two independent equations, for if we assume the validity of the equa
tion Ti~.rt = 71:721 it follows immediately, because of the identities ri + t« + t3 = 0
and 71 + it + 7a = 0 [see Eqs. (16-7) and (16-8) J, that the second equation

ti'.t*. = 7j:7j
holds.
230 THEORY OF FLOW AND FRACTURE OF SOLIDS

In words: principal rates of shear are proportional to


the three (he corre

sponding three principal shearing stresses.


Rule 3 containing also rule 2 of the constancy of the volume

ei + «2 + e3
= 0

is expressed in a third form by the three equations:

(16-11)
</>

in which an undetermined quantity since we substitute the values


is

if
of the rates of strain i\,ii,i3 given by the right sides of Eqs. (16-1 in the

1)
right sides of the proportions in Eqs. (16-9), we obtain the left sides of
these last equations and see that Eqs. (16-11) satisfy the proportions
(16-9), and we take the sum of the three Eqs. (16-11) we obtain
if

+ «j =
+

«i «2
0
or Eq. (16-2a).
Equations (16-9) or (16-10) contain law of distortion which might
a

generally be stated as follows for any direction in space:

= 3*r. (16-12)
^

According to rule in plastic mass slowly deforming the shearing stress


a
3

proportional to the corresponding rate of shear.


is
t

The stress-strain relations expressing Hooke's law for elastic bodies

*i = ki — "0(,2
+ <T3)]
^

" = ^2

""^
'
^
E

'

= ~
«s ^3 V0^<tx
E

(E modulus of elasticity, Poisson's ratio) have a form similar to Eqs.


p0

(16-11). We may state that for plastic mass in steady slow motion and
a

provided that the strains remain small Eqs. (16-11) express the stress-
rate-of-strain relations. They differ from those valid for an elastic
material in three respects: first, they contain instead of a constant
<t>

modulus of elasticity quantity which variable; second, Poisson's


E

is
a

1
RULES IN THEORIES OF FLOW 237

ratio co in them is equal to


^2, for incompressible material; and third,
instead of the components of strain ti,ti,t3 in Eqs. (16-11) the components
of the rates of the permanent strains <i,e2,€3 appear on their left sides.
Remembering that for steady slow plastic flow when the principal
stresses in every element of material do not change their magnitude and
their directions do not rotate relative to the material and when the
strains remain small we may substitute instead of the rates of strains
the strains themr;elves. Under the conditions just stated we may write

—2~y (16-14)
<t>

e3
= ^3
CHAPTER 17

TESTS ON YIELDING
AND FRACTURE UNDER COMBINED STRESS

The conditions under which various materials start to deform plas


tically have been the subject of many experimental investigations during
the last hundred years. Among these the tests of J. Guest1 with ductile
metals; of A. Foppl2 with rock materials; of T. v. Karman3 with marble
and sandstone under combined stress; of R. Boker4 with the same sub
stances and zinc; of F. B. Seely and W. I. Putnam6 with steel; of F. E.
Richart, A. Brandtzaeg, and R. L. Brown" with concrete under combined
stress; of the author and W. Lode7 with iron, copper, and nickel; of
P. Ludwik8 with steel and other metals; and of M. Ros and A. Eichinger'
with metals and rock materials; of Sir Geoffrey Taylor and H. Quinney10
on aluminum, copper, lead, cadmium, mild steel, and glass; of E. Siebel
and A. Maier11 on 0.37 and 0.60 per cent carbon steel and brass; of J. M.
Lessells and C. W. MacGregor12 on nickel-chrome-molybdenum steel; and
of D. J. McAdam,1' might be mentioned.14
I Phil. Mag., 1900.
* Mill. mech. tech. Laboratorium(Munchen), 1900.
*
Forschungshefl 118, and Z. Ver. dent. Ing., 1911.
4
Dissertation, Technische Hochschule, Aachen, 1914.
*
Univ. Illinois Eng. Expt. Sta. Bull. 115, vol. 17, 1919.
« Univ. Illinois
Eng. Expt. Sla., Bui. 185, vol. 26, 1928.
7 See author's
papers in Berichte des Werkstofiaussehuss, Verein deutscher Eisen-
htittenleute, Dttsseldorf, 1925; also Proc. 2d Intern. Congr. Applied Mechanics,
Zurich, 1926; and W. Lode, Mitt. u. Forschungsarb., Heft 303, 1928.
8 Bruchgefahr und Materialprtifung, Schweiz. Verband Material prufung. Tech.
Ber. 13, Zurich, November, 1928; cf. also his "Elernente der technologischen Mech-
anik," Verlag Julius Springer, Berlin, 1909; and other reports of his in Z. Ver. deut.
Ing. and in <StaW u. Eisen in recent years.
9 Proc. 2d Intern. Congr. Applied
Mechanics, Zurich, 1926, also Eidgendss. Material-
priifungsanstalt E.T.H. Ber. 28, Zurich.
10 The Plastic Distortion of
Metals, Trans. Roy. Soc. London, ser. A, vol. 230,
pp. 323-362, 1931.
II Einfluss mchrachsigcr Spannungszustaende auf das Formacnderungsvermoegen
metallischer Werkstoffe, Z. Ver. deut. Iny., vol. 77, p. 1345, 1933.
11 Combined Stress Experiments on a Nickel-Chrome-Molybdenum Steel, J .
Franklin Inst., vol. 230, p. 163, 1940; also Proc. 5th Intern. Congr. Applied Mechanics
Cambridge, Mass., 1938.
11McAdam and his associates at the National Bureau of Standards have investi
gated the flow and fracture of metals under combined stresses and reported the results
of their tests in numerous papers, most of which were published in Metals Technol.,
Proc. ASTM, and Proc. ASM.
14 A detailed discussion of these and other older experimental investigations
on the
238
TESTS UNDER COMBINED STRESS 239

17-1. Tests by Guest, Foppl, and Karman. J.


Guest carried on his
tests with thin tubes of steel, iron, and copper. These were subjected
either to pure axial tension, to axial tension simultaneously with internal
hydraulic pressure, or to a twisting moment and a tensile force. The
diameters of the largest principal (Mohr's) circles representing the state
of stress at the yield points were found to be equal for these ductile metals,
except for small differences which were especially noticeable for the case
of pure torsion (tn = c, <n = 0, a3 = — c). For these tests Guest con
cluded that the condition of yielding of the metals he had investigated
was expressed by a linear equation connecting a\ — <r3 and <j\ + <^3
(<ri > (T2 > ffj) [this equation was quoted in Chap. 15; see Eq. (15-23a)].
In the tests of A. Foppl cubes of rock were loaded in a special apparatus
upon two or four sides and tested in compression to failure. The result
was that the intermediate principal stress appeared to be without influ
ence on the breaking strength. Only tests to fracture were made.
Relative to the plasticity of materials, which under an ordinary tensile
or compression test behave in a brittle manner, the tests under axial and
simultaneous lateral pressure by T. v. Karman and R. Boker may be
mentioned. Both loaded cylindrical test pieces of marble and sandstone
either in axial compression or in a pressure container of steel under axial
compression combined with high lateral hydraulic pressure. The results
of their tests may be summarized as follows: With increasing hydraulic
pressure the diameter of the largest principal Mohr's circle increased and
approached a limit for the highest hydraulic pressures. The determination
of this maximum diameter proved difficult in these tests under the higher
pressures, since the marble possessed, under these circumstances and the
high pressures, a steep stress-strain curve having no definite yield point.
With respect to the influence of the intermediate principal stress no con
clusion could be drawn from these tests. The acute angles which the
slip planes made with the direction of the principal compression were
measured in the test pieces, however, and were also determined from the
Mohr limiting curve. In the enveloping curve of the principal stress
circles the angle which the normal of the limiting curve makes with the
a axis is the angle of the slip planes (Figs. 15-33 and 15-34). Because of
the large plastic deformation, the angle observed at the end of a test
had to be corrected to allow for the change in shape of the test piece. In

conditions of yielding not quoted here may be found in W. Lode's paper (loc. cit.).
Cf. also author's paper, Theories of Strength, Trans. ASME, 1932. Extensive investi
gations on the flow, the laws of strain hardening, and the conditions of fracture of
ductile metals which were carried out during the last war in the United States are
reported in Sec. 17-7.
240 THEORY OF FLOW AND FRACTURE OF SOLIDS

this way satisfactory agreement between the measured and calculated


a
angles of slip was obtained. These angles increased in plastic marble
from 53 deg with no hydraulic pressure to 73 deg with a hydraulic pres
sure of 685 atm. The tests showed that a low pressure corresponds to
behavior as a "brittle" material with a small angle of slip, while the large
pressure corresponds to behavior as a "plastic" material with an angle of
slip with approaches 90 deg.
If we plot the maximum differ
ence in principal stresses <T\ — <ri,
under constant external
pressure,
as ordinates, against axial com
pression as abscissas, we obtain
"deformation curves" under vari
ous external pressures. These are
represented in Fig. 17-1 for the
Karman tests on marble. In this
figure we see the various well-
1,,, known types of stress-strain curves
of brittle, of partly brittle, and of
ductile materials to appear in one
1 2 3 4 5 6 7 8^ 9 %IO
and the same
material. If the
OH IT COMPRESSION stress-strain curve of the marble
Fig. had a sharp break corresponding
17-1. Kiirmiin's
marble cylinders
tested under combined axial and lateral to a very definite yield point with
hydrostatic pressures. a\ axial compres
subsequent decrease in stress (see
sive stress, at lateral hydrostatic pressure.
The principal stress difference ax — at is curve for 235 atm pressure in Fig.
plotted against the axial unit compression,
17-1), very pronounced flow or slip
while in each test the lateral pressure at
was kept constant. lines were noted on the material.
Under the higher external pres
sures, the test pieces under compression exhibited a comparatively
more uniform bulging than under the lower external pressures, when
they bulged out only in the middle. After an ordinary compression
test, the microstructure of the material showed countless fine cracks and
fissures and the crystals appeared to have loosened along their boundaries.
On the other hand, if the marble was deformed plastically, "twin mark
ings" in the grains were especially numerous. The lateral hydrostatic
compression had prevented the breaking up of the marble crystals.
Although marble is brittle under ordinary conditions of loading, these
tests clearly show that under high constraining pressures it became con
siderably plastic. Consequently the calcite crystals composing marble
deformed by twinning as mentioned before.
The results of the tests by Karman which were reproduced in Fig. 17-1
TESTS UNDER COMBINED STRESS 241

referred to marble cylinders which were brought to the plastic state by


increasing the axial compression stress while the cylinders were exposed
to a constant high lateral pressure. R. Boker ran a second series of
similar tests with marble and cast zinc cylinders in which he held the axial
stress constant while the lateral pressure was gradually increased until
the cylinders started to deform plastically. Under these conditions the
cylinders necked down in their central portion. The hardened steel
plates between which the test cyl
inders were deformed in both groups I V M2

of tests prevented the ends of the


specimens from expanding or con
tracting freely because of the radial
frictional forces exerted on them.
The distortion of these materials
known to behave in a brittle man
ner under ordinary compression or
tension tests is indicated in Fig.
17-2; (a) and (b) show the barrel-
WM
shaped bulging of the specimens of
Ic)
Fio. 17-2. Shapes of plastically deformed
marble observed by Karman when solid cylinders (a) and (6) in marble
the axial compression stress was (Karman) when axial compression stress
was larger than lateral compression stresses,
larger than the lateral fluid pressure (c) in cast line (Boker), when lateral com
pression stresses were larger than the axial
and (c) the plastic necking observed
compression.
by Boker on a cast zinc cylinder
(a material which under simple tension behaves in a brittle manner) when
the lateral pressures were larger than the axial compression stress.
If the variable n which was introduced in Eq. (16-3) is used as a meas
ure of the intermediate principal stress aY and if <r\ > <rY > <t%, we may
state that in the series of tests in which the absolute value of the axial
compression stress <rY was larger than the absolute value of the lateral
pressure, i.e., when a\ = <r% = — p\, a3 = — p3, p3 > Pi the variable n
was equal to 1, and in the series of tests in which the lateral compression
stresses <T3 were the larger ones <r\ = —pi, <r2 = <rV = — p3, p3 > p\,
n was equal to —1.
The first series of tests (Karman, n = 1) may be also characterized as
simple uniaxial compression tests on which a high value of a state of tri-
axial equal pressures (<r/ + <rY + <rY)/3 = — (2pi + pi)/3 is superposed,
while the second series of tests (Boker, n = — 1) represents uniaxial
tension tests on which a high value of pure hydrostatic pressure

- (Pl + 2pa)
3
242 THEORY OF FLOW AND FRACTURE OF SOLIDS

is superposed. Boker and Karman found for marble that the major
principal stresses causing plastic distortion in both series of these tests
= —
1) in the graphical representation of their principal cir
(n = 1 and n
cles in a a,r plane had a common envelope in two curves which approached
two parallel straight lines r = +c when the mean pressure increased to
large values.1

In an additional remarkable series of tests Boker subjected solid marble


cylinders to combined torsion, an axial load in compression, and lateral
fluid pressure using the apparatus shown in Fig. 17-3. It is well known
that a round bar of a brittle material under torsion fractures along a
surface, which intersects the surface of the cylinder in a helix inclined at
1 These tests thus disprove the validity of the theory of strength referred to Cou
lomb according to which the inner resistance of solids against plastic distortion should
depend on an inner friction in a material. According to Coulomb's theory of internal
friction, a\ — a% — c\ + c2(<ri + a3), whereas the tests with marble led to the conclu
sion that the equation of the enveloping curve <r, — a, = f(a, + a3) represents curves
approaching the asymptotes r = +c = const (see also See. 15-13).
TESTS UNDER COMBINED STRESS 243

an angle of 45 deg with respect to the generatrices of the cylinder (Fig.


17-66). The helix is perpendicular to the direction in which one of the
two principal stresses in the surface of the twisted bar acts, namely, the
one which is a tensile stress. If the solid test cylinder is stressed in pure
torsion and simultaneously exposed to a state of high hydrostatic pressure,
the tensile principal stress will be reduced and the other principal (com
pression) stress increased. Under these conditions the torsional couple

Figs. 17-6. Bolter's tests with marble cylinders simultaneously stressed under lateral
fluid pressure, an axial load, and by a torsion moment, (a) Specimen undeformed. (6)
Simple torsion fracture by cleavage, (c) First loaded by ar = at = —580 atm, and axially
by a, — 2,395 atm, then under these stresses twisted to fracture. The steep helix is the
cleavage crack, inclined at 5.3°; the other helix, inclined at 8° is a shear fracture. Note
slight lateral plastic bulging of specimen, (d) Loaded by equal stresses a, — at = ax. then
twisted. The helix (cleavage crack) went around a full turn of 360°. (c) First loaded by
ar = at — —665 atm and axially by a, = —2,770 atm, then under these stresses twisted to
fracture. The steep helix is the cleavage crack, inclined at 62°. The shear fracture can
barely be seen at an angle of 18°. (The short additional cracks branching off the main
cleavage crack are secondary fractures.) Note the considerable lateral plastic bulging
and the permanent twist in the test specimen.

can be increased beyond its former value and the cylinder of a brittle
material can be deformed plastically.1 By superposing subsequently an
additional axial compression load Boker was able to increase the angle
of the helical cleavage crack with respect to the planes perpendicular to
the axis of the cylinders along which the marble specimens broke in ten
sion. Two broken marble cylinders which were tested in this manner
are shown in Fig. 17-6c, e. In a few cases the specimens disclosed
evidences of both types of fracture surfaces, of a cleavage and of a shear
fracture. Both appeared as parts of helices on the surface of the solid
1 All specimens
exposed to high hydrostatic pressures on their cylindrical surfaces
were covered by a very thin metal foil which prevented the pressure fluid from pene
trating in the crevices of the marble.
244 THEORY OF FLOW AND FRACTURE OF SOLIDS

cylinders, the former as a steeplyinclined helix, the latter as a helix


inclined at an angle of 45 deg relative to the steeply inclined tension
crack which one would expect if the second fracture was of the shear type.
The views of the broken cylinders clearly show the signs of permanent
twist and of plastic compression.1
The experiments by Karman and Boker were repeated at Harvard
University by David T. Griggs2 with a few rocks, marble, and limestone
in types of tests practically identical with those the former two investi
gators had made. Griggs, in contrast to the work of Karman and Boker,
exposed most of the cylindrical specimens of very porous rocks such as
limestone which he tested in their surface to the action of the pressure
fluid (kerosene) which was permitted to penetrate in the crevices of the
material under the high hydrostatic pressures. His test results were
therefore obscured through this secondary effect.
17-2. The Influence of the Intermediate Principal Stress on the Yield
ing of Metals. Regarding this question W. Lode,3 at the suggestion of
the author, has carried out a number of tests at the Institute for Applied
Mechanics at the University of Gottingen (Germany). In thin-walled
tubes of iron, copper, and nickel subject to axial tension combined with
an internal hydraulic pressure, it was found possible to produce three
states of stress in the region of tensile stresses, which for the sake of
brevity will be designated by I, II, and III
and which in a certain sense
could be considered to correspond to the simple cases of pure tension,
pure compression, and pure shear. Mohr's principal stress circles of
these three states of stress are shown in Fig. 17-7. Since the intermediate
principal stress <r2 in the stress distribution I was equal to the small
est principal stress, in the stress distribution III
it was equal to the largest
principal stress, and in the stress distribution II equal to half the largest
principal stress (in all three cases the smallest principal stress being
approximately or exactly equal to zero), the influence of the intermediate
principal stress on the value of the diameter of the largest principal circle
should be quite marked in these tests. Moreover, the simultaneous
small change of the mean tension (<r\ + <tj + <r3)/3 could be neglected
since its influence on the characteristics of ductile materials is small.
With one test piece a number of stress-strain curves were taken. In
plotting stress-strain diagrams one of the three principal extensions (e.g.,

1 Some cracks visible on the surfaces of the broken


specimens in Fig. 17-6 are of
secondary nature and were formed after the cleavage and shear fractures split the
specimens.
2 Deformation of Rocks under High Confining Pressures, J. Oeol., vol. 44, No. 5,
p. 541, August, 1936.
1 Loc. cit.
TESTS UNDER COMBINED STRESS 245

H./i„0 m.u„i

„ 0,
02

Fig. 17-7. Principal stress circles for Lode's three states of stress, I, II, III, produced in
thin-walled tubes subjected to combined axial tension and internal hydraulic pressure.
I. For tube under axiai tension alone. II. For tube under internal pressure alone. III.
For tube under combined action of both.

6 7
EXTENSION mm

I TESTSNO 12 3 4 5 6 7 8 9 10

\ H = -1.00-1.00 '0.02*002 -100*0.96-1.00-1.000.00 -1.00


(TESTSNO. II 12 13 14 15 16 17 .18 19 20
\ -0.43-069 -100*0 35*0.62 -1.00 0.00*,.97-1.00 -100
Flo. 17-8. Lode's tests with steel tube under various combined tensile stresses. Abscis
sas: axial extension of 100 mm gauge length in millimeters. Ordinates: greatest principal
stress differences in kilograms per square millimeter.

the extension of the specimen along the axes) was taken as strain and the
difference of the largest and smallest principal stress as stress. Curves
plotted in this manner are shown in Fig. The disturbing influence
17-8.
of the gradual strain hardening of the metal, due to progressive plastic
deformation (work hardening), was eliminated in this way. The results
of various tests are represented in Figs. 17-8, 17-9, and 17-10. In Figs.
17-9 and 17-10 the abscissa n denotes the quantity which was introduced
in Eq. (16-3) as a measure of the intermediate principal stress <rj:

a -
M
=
(<T,
- + ,r,)/2]
f(<r,

<r,)/2
(17-1)
246 THEORY OF FLOW AND FRACTURE OF SOLIDS

As already stated, the quantity n varies between the values — 1 and +1.
The ordinates in Fig. 17-9 represent the ratios of the stress difference

<ri <r3 required to initiate flow to the yield stress for simple tension <r0.

We see that case I, n = — 1, <r2 = <T3 = 0, corresponds to a state of


simple tension in axial direction in the tube; furthermore case II (a tube
having both ends closed subjected to pure internal hydrostatic pressure)
corresponds to a value of n = 0,
<r2
= <ri/2, <r3 approximately equal
to zero, or to "a state of pure
shear," and case III, in whicb the
intermediate principal stress <t. was
-1.0-0.8 -0.4 0 *0.4 * 0.8 * 1.0

Fio. 17-9. Variation of the greatest prin made equal to the major principal
cipal stress difference with intermediate stress <ti, while a3 again was approx
principal stress according to Lode's tests.
imately equal to zero (a tube sub
jected to internal pressure and an axial tension of the required amount)
corresponds to "a state of simple compression," 11 = 1.
The observed points in Fig. 17-9 are grouped along a symmetric curve
of parabolic shape passing through the points m = — 1 and n = 1 ,
(<r1

a3)/<r0
= 1. The curve connecting these two points shown in
Fig. 17-9 represents the values of
the stress ratio (<r1 — <r3)/<ro com »l.0
V
puted from the condition of flow *0.8 *

root = const. This curve reaches *o.e


its vertex at *0.4 *

— = 1.155. *0.2
n = 0, (<r1 <r3) /<r0 1 :
0 A
-
The agreement between the test -O.B
points and this curve in Fig. 17-9 -0.4
is very satisfactory. Figure 17-10 -0.6 \—
reproduces the observed values of
-0.8%
Eq. (16-6), page 234]
- -
[see

- 2t2

41 43
(17-2)
-l.0\ 0.8
Fio.
-0.4
17-10. m and v values
ti «3

in their dependence on n, again showing that the relation v = 11 postulated


for the third rule of flow is verified, although a systematic deviation of the
points from the inclined straight line v = n is noticeable.
By combining the results of the tests made with hollow cylinders
stressed by an axial tension load and by internal hydraulic pressure
(Fig. 17-11) with tests (using
a few torsion and combined tension-torsion
the machine shown in Fig. 17-121 and hollow specimens) W. Lode found
1 Designed by v. A.
Bohuczewitz and Sonntag by the Losenhausen Works,
Dusseldorf, in 1925.
TESTS UNDER COMBINED STRESS 247

Fig. 17-11. Amsler high-pressure pump and Losenhausen testing machine for combined
stress test.

Fio. 17-12. Losenhausen tension-torsion machine.


248 THEORY OF FLOW AND FRACTURE OF SOLIDS

no appreciable effect of the mean tensile stress <r = (a\ + <r2 + <r2)/3 on
the octahedral shearing stress t0 required for initiating yielding of the
metals he had investigated.
17-3. Tension -Torsion Tests by Taylor and Quinney.1 The first named
author has extended his most careful observations on the plastic deforma
tion of single metallic crystals2 to polycrystalline ductile metals. By
measuring the change of the internal volume of the tube during the
strain by filling it with water and reading the movement of a column in a
capillary tube, he and Quinney were able to determine also the relation

(AXIAL STRESS ff) ~ —*• (AXIAL STRESS CT)

Fio. 17-13. Fio. 17-14.


Figs. 17-13 and 17-14. Combined tension-torsion tests by G. I. Taylor and H. Quinney
for copper, aluminum, and mild steel.

existing between the quantities n and v in a more perfect way than was
the case in Lode's early tests. The volume of the bore of a tube stretched
plastically in axial tension should not change. The same is true for pure
torsion and hence also for all combinations of both, provided that n = r.
Hence the possible changes of volume of the water inside the bore could
be utilized to find any observable deviations from the law n = v. These
tests led to the results reproduced in Figs. 17-13 to 17-15. According to
the plasticity condition t„x. = const, yielding in a tube subjected to
tension-torsion should start when

<r2 + 3t2 = <r02


= const
(17-3)

where a is the axial tensile and r the shearing stress, and <r0 the yield stress
for pure tension. However, according to the theory of maximum shear
(Guest-Mohr), yielding would start as soon as

a* + 4t* = <r0* . (17-4)

Plotting a and r as rectangular coordinates, the expressions above give


ellipses with the semiaxcs <r0, <r0/ "\/3 and <r0, <r0/2, respectively (Fig. 17-13).
Taylor, and II. Quinney, The Plastic Distortion
I.,

1 G. of Metals, Trans. Roy.


A,

Soc. London, ser. vol. 230, pp. 323-362, November, 1931.


See Chap.
2

7.
TESTS UNDER COMBINED STRESS 249

Figures 17-13 and 17-14 taken from Taylor and Quinney's report show
that for aluminum or copper the test points very closely follow the upper
ellipse corresponding to Eq. (17-3), disproving Eq. (17-4). For mild
steel, however, the agreement between the observations and the curves
was by no means as good. All points were found lying above the von
Mises ellipse. In any case they were closer to this than to the other
ellipse. Since the observations of the volume filled with water showed
anomalies which appeared closely connected with the breaks of the stress-
strain curve, most probably the different behavior of mild steel in these

Fio. 17-15. v = f(ji). (According to testa by G. I. Taylor and H. Quinney.)

tests must be attributed to the nqnuniformity of the first yielding over


the length of the specimen (yielding in thin layers) due to the phenomenon
of upper and lower yield points.
Figure 17-15 shows the test results of Taylor and Quinney with regard
to the kinematics of the plastic flow. The scattering of the points in
Lode's n,v diagram appears to be eliminated here, and for the relation
v = f(n) very distinct curves were obtained, although not very different

from the straight line v — ix.


One interesting result was that the curves for the hexagonal metal
cadmium and for lead approached the straight line n = v much more
closely than did those for metals (copper, aluminum) with a regular
lattice. The points for the amorphous heated glass were almost exactly
on the line v = as was to be expected for viscous flow of a substance.
It seems probable that the type of lattice structure within the individual
grains has something to do with the shape of the v = f(n) curve. This
250 THEORY OF FLOW AND FRACTURE OF SOLIDS

was suspected by Lode, who tried to explain the shape of the actual curve
v = f(n) by the statistical effect of a great number of grains oriented at

random with their axes, but yielding according to the well-known modes
of the single metallic crystals.
Summing up the results of Lode's and of Taylor and Quinney's tests
we can say that the intermediate principal stress o-2, contrary to what
should be the case corresponding to the Mohr or to the maximum shear
ing stress theory for the investigated metals, has a marked influence on
the diameter of the major principal
lA
/*
stress circle. According to the con
AXIAL
STRESS ♦7 ^»>^ dition 7W, = const for the state of a
pure shear = the diameter of

0)
(fi
Oi
this circle should be equal to <ra,
IAS n =
crt/cra
whereas the tests clearly indicated
t

that the latter was considerably larger


0, AXIAL STRAIN €a ^- and near the value 2<r0/\/3 = 1.155ct0
Ludwik's apparent de predicted by the theory rMl = const.
crease of ductility of a metal tested
We conclude from the test results in
under biaxial stresses.
Sees. 17-2 and 17-3 first, that the con
dition of yielding for the ductile metals within experimental errors, well
is

expressed by the equation

V2 (To
Toot = VVl _ <Tl)'1 + (<72 TlY + (<73

Cl)2 = = const
(17-5)

,
^

second, that the third rule of plastic flow in the metals = sufficiently is
v

fi

verified to warrant its introduction in the theories of flow, in spite of the


observed deviation of the curves connecting the experimental points of

v
and from the straight line = n; and third, that = for amorphous
v
n

solids at high temperatures),


(glass thus proving that glass at high
temperatures flows like perfectly viscous substance.
a

17-4. Ludwik's Remark Concerning the Ductility of Metals under


Biaxial States of Stress. Suppose that stress-strain curves are recorded
for a ductile metal by testing tubular specimens stressed by an axial load
and an internal pressure under biaxial states of stress, e.g., by plotting
the true axial stress <r„ as a function of the axial strains «„• Suppose that
aa > <rt and that the ratio of the tangential to the axial stress n = at/<ra
kept constant during each test. For simple (axial) tension = a
0)
(n
is

curve like OR0 (Fig. 17-16) would be observed. When the ratio n is
increased, shorter curves OR\, OR2, . . . would be recorded, the points
Ro, Ri, Ri, . . . representing the points at which the fractures occurred.
The decrease of the axial strain e„ at which these curves terminate at frac
TESTS UNDER COMBINED STRESS 251

ture with the increasing stress ratios n was interpreted by P. Ludwik,1


who drew a diagram like Fig. 17-16 as a decrease of the ductility of the
metal with increase of the ratio <r</<r0 of the two tensile principal stresses ct
and aa. A comparison of such stress-strain curves, however, does not
convey a satisfactory method for expressing the ductility of a metal
under biaxial stressing, since neither the axial stress ca (in the presence of a
circumferential normal stress <rt), nor the axial strain «,, is a satisfactory
variable measuring the intensities of stress and strain under a biaxial
state of stress. Shearing stresses and unit shears must be proposed
instead of <ra or e„ which depend on all three principal values of stress and
permanent strain if the correct variables expressing the intensities of
stress and strain are to be established to measure "the ductility" of a
cold worked metal.
17-5. Further Tests. The results of combined stress tests by Lessells
and MacGregor made with thin-walled tubes of a nickel-chrome-molyb
denum steel also agreed favorably with the flow condition root = const.
In their series of tests a few cases were included in which one of the
principal stresses was a compression stress. Ros and Eichinger arrived
at similar conclusions from their series of tests under combined stress
with various materials.
17-6. Plastic Collapse of Deep-well Steel Casing under External Pres
sure and Axial Tension. The heavy-walled seamless steel tubes used in
deep oil wells have to withstand large external pressures exerted on the out
side of the tubes through the weight of the surrounding overlying strata of
rocks.2 The American Petroleum Institute dealt with the problem of the
"plastic" collapse endangering steel casing in these great depths in a
symposium in 1939. 3
It was shown jointly with J. H. Holmquist4 that, as for the plastic
buckling of short straight steel bars (columns) under axial compression,
the danger of plastic collapse through the circumferential compression
stresses in the heavy-walled steel casing limits its application under the
external pressure. When tubes must be moved in the well and pulled
out against the frictional resistance exerted by the surrounding rock or
sand strata in addition axial tensile stresses may be present in the tube
i Z. Ver. deut. Ing., vol. 71, p. 1532, 1927.
* Oil wells having depths of 2 miles or more have been drilled; in a depth of 2 miles
= 3.2 km the weight of the overlying rocks produces a pressure of the order of 900 to
1,000 atm or 12,780 to 14,200 lb/in.s.
3 American Petroleum Institute Standards on Setting Depths and Collapsing Pres

sure, forum during 1939.


4 "Theoretical and Experimental Approach to the Problems of Collapse of Deep
Well Casing," American Petroleum Institute (Drilling and Produeting Practice),
1939 meeting, Chicago, pp. 392-420.
252 THEORY OF FLOW AND FRACTURE OF SOLIDS

walls. If the steel has a better defined yield point in a pure uniaxial
compression test, when the condition of plasticity is reached under the
combined forces, this determines the critical combination of the stresses
for collapse. The effect of added axial stress, particularly tension, on
plastic collapse of tubing was determined independently in a series of
collapse tests by S. H. Edwards and C. P. Miller1 using different grades
of steels having better defined yield points ot>. By plotting <ra/oo and
ffi/o-o all test data could be assembled in one figure (Fig. 17-17), regardless

/
\\
^1.00 -.75 -.50 -.25 .25 .50 .75 1.00

/
Q

/
/ \ /
.25

/ / .50
K
\c'

/
A
.75

jj*
/ I
i
- "^
e
I.OO%

1.25
a,
• TENSION COLLAPSE
Oc -AXIAL STRESS (TENS. S COMP)
01
» TENSION STRETCH a, -HOOP STRESS fCOMR)
• COMPRESSION COLLAPSE Co 'YIELD STRESS
Fig. 17-17. Fit of Edward's and Miller's tests to ellipse of plasticity.

of the fact that they were obtained on different grades of steel. The
curve shown in Fig. 17-17 represents the lower half of the plasticity ellipse
obtained from Eq. (17-5) by assuming <ro,<r,,0 as the principal stresses:

cra2 Oa<*t + fi2 (TO* (17-6)

With the exception of one single point observation points are well
grouped near the ellipse.
17-7. Tests on the Strain Hardening and Fracture of Metals under
Combined Stresses. P. Ludwik was one of the first technologists to
propose a general strain hardening function for the ductile metals.2 He
1 "Effect of Combined Longitudinal Loading and External Pressure on the Strength
of Oil-well Casing," American Petroleum Institute (Drilling and Production Prac
tice), 1939 meeting, Chicago, pp. 483-502.
* "Elemente der technologischen Mechanik," Verlag Julius Springer, Berlin, 1909,
57 pp. Also Ludwik, P., and It. Sohbu, Stahl u. Eisen, vol. 45, pp. 11, 373, 1925
(comparative tension, compression, and torsion tests on copper).
TESTS UNDER COMBINED STRESS 253

attempted to construct this function from tension, compression, and


torsion tests by assuming that in the metals in which the internal shearing
resistance against a permanent distortion was independent of the normal
stresses in the planes of slip the maximum shearing stress t™ could be
expressed as a function of the true maximum shearing strains -yw. After
introducing the logarithmic strains i = In (1 + e) he also denned a new
measure for the increments of the maximum unit shears which he referred
to the instantaneous configuration of a specimen (instead of to the orig
inal unstrained shape of it). The curve rm = /(7m»«) he called the "flow
curve" (Fliesskurve) of the metal. He believed that his tests with a few
metals supported his assumptions.1 He was also one of the first investi
gators who noticed that the flow stresses of metals of low melting points
increase with the logarithms of the strain rates.
The strain hardening of copper and of a low-carbon steel under com
bined tension and torsion was investigated by R. Schmidt2 who simul
taneously twisted and pulled permanently hollow thick-walled cylinders
in the tension-torsion machine mentioned in Sec. 17-2 and maintained
the ratio of the shearing and the axial normal stress at a constant value
during the permanent distortion of the specimens. From his tests
Schmidt concluded that a strain hardening function Toct = /(7oot) expresses
the behavior of a ductile metal under increasing values of the stress.*
Before discussing further tests it may be in order to state briefly some
mechanical facts on which a theory of the strain hardening of metals has
to be based:
1. The six components of permanent strain must express finite dis
tortions of an order of magnitude similar to the ones which are observed
just before the metal breaks.
2. Isotropic flow may be assumed within a moderate range of strains
observed preceding the fracture.
3. The volume of the metal during this deformation does not change.
4. According to the tests by Lode and by Taylor and Quinney described
in Sees. 17-2 and 17-3, the rates of shear with which the planes in an

1 The maximum value of the true shearing strains fmax (amounting to the natural
shears of the first kind, defined in Chap. 13, p. 142, Eq. (13-34) for a tension (or a com
pression) test Ludwik took equal to fm** = 2f i, which unfortunately is not correct
and should have been fm*x = 3«i/2, i\ being the natural strain in axial direction.
• Dissertation, Gottingen, 1932, 23 pp. Also Ing. Arch., vol. 3, 1932.
3 The value of his comparisons between various other plausible relations connecting

plastic strain with stress is obscured through the fact that he assumed as the condi
tion of incompressibility «i + e2 + e> = 0 which is not valid for finite strains of the
order of 30 per cent which he applied in his tests and that his strain variable ^oot was
not defined correctly for strains of this order of magnitude.
254 THEORY OF FLOW AND FRACTURE OF SOLIDS

element of the metal instantaneously slip relative to each other are pro
portional to the shearing stresses acting in these planes.
■ 5. The variable has to be defined adequate to measure the intensity
of strain (cold deformation) under the most general states of strain, and
a corresponding intensity of stress should be found.
In order to obtain information on the validity of some of these assump
tions and particularly in an attempt to define the two last-mentioned
variables within a range of strains of finite, moderately large magnitude,
two series of tests were carried out at the Westinghouse Research Lab-

Fia. 17-18. Hollow cylindrical tost specimen.

oratories by E. A. Davis,1 one with copper and the other with a soft grade
of steel, in which these metals were deformed slowly under states of biaxial
stresses while the ratio of two of the principal stresses was kept approxi
mately constant, the third principal stress remaining small.
Test specimens in the shape of hollow cylinders of annealed copper
have been subjected to combined axial tension and internal fluid pressure,
and their distortion carefully recorded. Thirteen specimens of the shape
reproduced in Fig. 17-18 were tested in two lots (A and B) in a 30-ton
tensile (hydraulic) machine and by using a three-piston high-pressure
pump of 1,200 atm, both constructed by Amsler & Company, Schaff-
hausen (Switzerland). By means of a suitable control synchronizing the
required opening of the inlet valves in the oil lines leading to the piston
of the testing machine and to the interior of the hollow test specimens a
constant ratio n = ot/ca of the circumferential to the axial stress could
be maintained during each test. The results of the copper tests are
reproduced in the Figs. 17-19 and 17-20. The abscissas of the observa
tion points in Fig. 17-19 are the absolute values of the maximum principal
natural unit shear, and the ordinates are the absolute values of the true
1 J. Applied Mechanics, December, 1943, and March, 1945.
TESTS UNDER COMBINED STRESS 255

maximum principal shearing stresses. The ratios n = <rJaa = % and


n = 2 correspond to a state of finite pure shear. No perceptible change
in diameter in the first and of axial length in the second case1 (tube under
pure pressure alone) could be observed in these two cases, in accordance
with the plastic stress-strain relations which were postulated2

li-*^!-^^), • - • • (17-7)

«i,«2,e3 denote the natural permanent principal strains. In Fig. 17-20

Fio. 17-18a. Thirty-ton testing machine with pendulum manometer coupled with high-
pressure pump of 1.200 atm (Amaler & Co., Schaffhauaen) for testing hollow cylindrical
specimens under biaxial stresses. (W eatinghouae Reacarch Laboratories.)

the same observations are reproduced in a plot of the true octahedral


shearing stress t0 as a function of the octahedral unit shear defined by
V
(ij

7o = %
— — — The test results can well
h)'1 +
+

(«i h)2 («3 «02.


be represented by one single curve.
It presumed that a theory of the flow of metals in the strain harden
is

ing range when the strains increase to magnitudes much larger than those
This has been also demonstrated in "long-time creep test" by F. H. Norton
1

with a thin-walled metal tube tested under internal gas pressure over a period of
4,000 hr at 1050T (Fig. 17-21). Since in thin-walled tube <r„ = <r,/2 in this case
a

according to Eq. (17-7) assuming <ri = aa, <r2 = <ro/2, <r3 = the permanent axial
0

strain — — tube with closed ends when deformed under internal pressure
0.

A
i.
ii

does not change its length in the plastic range of strains.


These stress-strain relations for a finite distortion are developed in Eq. (28-19).

250 THEORY OF FLOW AND FRACTURE OF SOLIDS

20,00a

18,000

LOT A '/• LOT B


a^<
l#?A

.>" #
X
£ J>
ort_

t,000-£

2,000

'0 .04 .08 .12 .16 .20 .24 .28 .32 .36 .40 .44

MAXIMUM SHEAR STRAIN


Fiq. 17-19.

LOT A' LOT B'


°J, •
*4

S 8,000

%
6,000

4,000

2,000

0 .04 .08 .12 .16 .20 24 .28 .32 36 .40 44


OCTAHEDRAL SHEAR STRAIN
Fiq. 17-20.
Figs. 17-19 and 17-20. Results of combined stress tests on copper. The circular dots,
triangular and square markings, etc., refer to tests in which the ratio n •* <r</<r0was kept
constant. The small circles in lot A represent the points of a tension test (n = 0) made
with a solid round bar, in lot B these points represent a pure tension test made with i
hollow cylinder. (Tests by E. A. Davis.)
TESTS UNDER COMBINED STRESS 257

which could be observed under these states of stress with positive (tensile)
principal stresses would have to take account of the anisotropy of the
metal which is known to develop when the grains are drawn out to longi
tudinal shapes,1 e.g, after sequences of continuous cold rolling or drawing.

.004

{LONGITUDINAL
1,200 IfiOO 2,000 2,400 2,800 3,200 4,000
TIME IN HOURS
Fig. 17-21. Long-time creep test of a hollow metal cylinder with both ends closed sub
jected to an internal gas pressure maintained at a constant value over a period of 4,000
hours (under 1,617 lb/inS pressure at 1,50°F). (Test by F. H. Norton.) No appreciable
creep strain in the axial direction could be observed (see lower curve representing the
longitudinal extension of tube) thus proving that a tube with closed ends when deformed
permanently by internal pressure does not change its length.

A similarseries of tests was carried out by Davis for the David Taylor
Model Basin with hollow specimens of a 0.23% steel that had been
annealed at 925°C. In order to test the tubes also under pure circumfer
ential tension (n = oo) the arrangement shown in Figs. 17-22 and 17-23

u- packing - SPECIMEN GRIP-

%-IO THREADS LEAD GASKETS -


Fio. 17-22. Cross section of arrangement to test stress ratio n = ai/oa — °o.

was used in which the axial load was carried by the central pin. Five
of the hollow steel cylinders after fracture are reproduced in Fig. 17-24
in the order from left to right for the cases n = 0 (simple axial tension),
1 For attempts in this direction the reader is referred to a recent
report by J. E.
Dobn, "Stress-strain Relations for the Plastic Deformation of Anisotropic Metals,"
report submitted to the David Taylor Model Basin, Navy Department, Washington,
D.C., June, 1947, 36 pp. (University of California, Berkeley) and to a paper by R.
Hill, A Theory of the Yielding and Plastic Flow of Anisotropic Metals, Proc. Roy.
Soc. London, ser. A, vol. 193, pp. 281-297, 1948.
258 THEORY OF FLOW AND FRACTURE OF SOLIDS

0.50, 1.00, 2 (pure internal pressure), and n = °c (pure circumferential


tension). When n = ai/<Ta = 1, <r< = <r<ione would expect that the frac
ture would proceed in any direction. The tube, however, broke along
an axial crack, as for the ratios « (see the enlarged views
n = 2 and n =
of the fractures in Fig. 17-25). It seemed of
considerable interest to determine that ratio n
at which the fracture changes from an axial
orientation to a circumferential fracture. The
photographic views of the fractured regions of
three more tubes are reproduced in Fig. 17-26.
These belonged to a series of five tubes which
were broken at the stress ratios

n = 0.75, 0.76, 0.78, 0.80, and 0.88,

respectively, showing that the fracture surface


changed from an axial to the circumferential
direction when n = 0.76. All fractured sur
faces were of the shear type, inclined at an
angle of approximately 45 deg with respect to
the tangent planes of the cylinders. The true
maximum shearing stresses at fracture were
reported in Sec. 15-13e, disclosing approxi
mately constant values for either of the two
orientations of the shear surfaces. The strain
hardening of this medium-carbon steel under
biaxial stresses over the range

is illustrated in Figs. 17-27 and 17-28. In the


Fio. 17-23. Arrangement to former again t„, = /("?m.«) and in the latter
test stresB ratio n = <ri/o-«
7<x-t
=
/(7<«t)The smooth curve
is plotted.1
in both figures connects the points of the ten
sion test n = 0. It can be seen that the observed points fit excellently a
common curve txi = /(foot) representing the true octahedral shearing
stress as a function of the natural octahedral shears foot.
The energy of local plastic distortion per unit of volume to deform a
small element of steel to the point at which it fails through the static

1 The points for those test specimens which broke in circumferential direction are
assembled in the left, those for tubes that broke along an axial crack in the right group
of points. R indicates the fracture point in each test.
TESTS UNDER COMBINED STRESS 259

shear fracture, which is

is found in the area under the general strain hardening curve times
the factor It decreases with the increasing ratios n to a con-

Fig. 17-24. Five hollow cylinders of 0.23% C steel tested by E. A. Davis under biaxial
stresses. Stress ratios n = <ri/oa — 0, 0.5, 1, 2, °° in order from left to right. At n =2
(tube with both ends closed under pure hydrostatic pressure) the specimen has not changed
its length. This tube (the fourth tube from the left ) represents the original length of all
specimens. Note that when n = =o the specimen has contracted in the axial direction.

siderable extent. (W = 66,000 in. -lb/in. 3 when n = 0, tension test;

W - 16,140 when n = 1; W = 5,390, n = 2; but W = 9,570 when


n = oo
.) It may be of interest to note that this apparent decrease
of local ductility is caused through an instability of the uniform mode
of deformation, which gradually develops also under biaxial states of
stress as in a simple tension test. Fractures were preceded by a short
period of local bulging or necking combined with a local decrease of wall
200 THEORY OF FLOW AND FRACTURE OF SOLIDS

Fro. 17-25. Fig. 17-26.


Figs. 17-25 and 17-26. Fractures of hollow steel cylinders.

.00 .10 .20 .30 .40 .50 .60 .70 .80 .90 .100

ifMAX.
NAZ MAX- SHEAR

Fio. 17-27.

/i . I i i i i i I i ; |
O .10 .20 .30 .40 .50 .60 .70 .80 .90 .100 .110 .120 .130
O .10 .20 .30 .40 .50 .60 .70 .80 .90 .100
-fQ

NAT. OCT. SHEAR


Fig. 17-28.
Figs. 17-27 and 17-28. 0.23% steel. (Tests by E. A. Davis on the law of strain hardening
under biaxial stresses.)
TESTS UNDER COMBINED STRESS 261

thickness caused when either the total axial load or the internal pressure
reached a maximum value.1
The amount of stored energy in the pressure fluid (oil) has a pro
nounced influence on the formation

L
A

5
and type of fracture surface.
few additional tests of this kind on
the same steel are reproduced in
the photographs of broken cylin
ders (Figs. 17-29 to 17-32). The
pressure energy available to accel
erate the spreading of a shear frac
ture that has just started to form
was increased by connecting the

fl
test specimens with an additional
large
served
voir.
pressure container which
as a pressure energy reser
In the tests in which the Fio. 17-29.
Figs. 17-29 and
fl
Fig. 17-30.
cylinders were con 17-30. Bursting tests with
hollow test
small-sized steel tubes (inside diameter
nected with the pressure energy 0.75 in., wall thickness ,.0625 in.). Left:
low-energy test. Right: high-energy test.
storage tank after short shear frac
Stress ratio: n = 2.
tures had first formed the crack
suddenly changed into a cleavage (herringbone) crack which was per
pendicular to the tangent plane of the cylindrical test specimens (see also
Sec. 15-6).
The preceding series of tests initiated a group of further investigations
on the flow and fracture of steel under biaxial stresses.2 From a semi-
1 The condition under which this instability occurs in a thin-walled cylinder of
ductile metal under biaxial stresses and local necking or bulging must start has been
investigated by W. T. Lankford and E. Saibel (Metals Technol., August, 1947).
This can be done by determining the condition under which the axial load or the
internal pressure in function of the corresponding plastic strain (axial or circum
ferential strain, respectively) becomes an analytic maximum. For this purpose these
authors assumed that the strain hardening function & = in pure tension (? true
stress, i natural strain) is approximated by a power function a = <•(«)". See also
Sachs, G., G. Espey, and G. B. Kasik, Trans. ASME, vol. 68, p. 161, 1946. The
condition when a local distortion in a thin circular diaphragm of a ductile metal
starts to develop during a bulging test under lateral pressure has also been investi
gated by the first named authors and in a paper by W. F. Brown and G. Sachs (1947
ASME meeting).
* These tests were carried out during and after the last war under a cooperative

plan at the Illinois Institute of Technology, Chicago, and at the University of Cali
fornia. Berkeley, under the joint supervision of the Research Committee for Plastic
Flow of Metals of the ASME, the National Defense Research Committee (War
Metallurgy Division) and the David Taylor Model Basin of the Navy Department
262 THEORY OF FLOW AND FRACTURE OF SOLIDS

killed medium-carbon steel (0.23% C, 0.47% Mn, 0.011% P, 0.042% S,


0.02% Si) similar to ship-plate steel, round bars were rolled at the works of
the Carnegie Illinois Steel Company in Pittsburgh from the same ingot.
From the round bars 24 cylindrical hollow specimens were machined of
3 in. inner diameter and % and Y± in. thickness; with a gauge length of

Fig. 17-31. Fig. 17-32.


Figs. and 17-32.
17-31 High-energy bursting tests with intermediate size steel tubes
(inside diameter 1.5 in., wall thickness 0.125 in.). Stress ratio : n = °o (left), n = 2 (right).
(Tests by E. A. Davis.)

19 in. and reinforced heads in shapes geometrically similar to the tube


reproduced in Fig. 17-18 and tested by G. K. Morikawa and LeVan
Griffis at the Illinois Institute of Technology, Chicago, under combined
internal pressure and axial tension.1
In of these specimens two slots in diametrically opposite positions were cut in
13
axial direction and rewelded; the remaining 11 specimens did not contain any welds.
Six of the unwelded tubes were tested in the original hot rolled condition of the steel,
the rest in annealed condition. Only the results of the tests of the unwelded speci
mens are reported here as given in Table 17-1. This grade of a low-carbon steel was
less ductile than the steel of a similar composition tested by Davis. The two tulies
tested in simple axial tension (n = 0) broke at an octahedral shear ft, = 0.941 and

in Washington, D.C., during the years 1943 to 1947. Cf. the papers presented in the
1948 Symposium on Flow and Fracture of Metals in /. Applied Mechanics, Septem
ber, 1948.
1 "The
Behavior of Steel under Conditions of Multiaxial Stress and the Effect of
Metallographic Structure and Chemical Composition on this Behavior," National
Defense Research Committee, Division 18, reports of February and July, 1945.
Table ,7ss. Biaxial Te,t, by Morikawa and Griffi, with Hollow Machined Cylinder, of ,.0% C Steel

\i
(Gaug, l,gth ,0 i,., i,,,, dia0,t,, -., in., wa0l thick,,,, i,.)

F,ac T,u, ,t,s, F,actu, data: Max Octah,d,al


Max ,xial Max
1

St,,, ,at,al ,t,ai,


Sp,ci Max tu, ,h,a,
ratio p,, load at Ta,g, u,it
0, axial p,, st,s,
"t ,u,, f,actu,, ,xial (To, tial at, ,h,a, St,s, to, Sh,a,
No. n 5, 5, lb su,
'

load, Ta,g,,
7

aa lb/i,.! lb. 0b/i,.1 lb/i,.J ,xia0 «« lb/i,. 1°

it

lb
/i,." tial lb/i,.>
-< 5,

,0 0.-, 0,,, ,7, 79.,, -,.,55 00 -0.0, ,.., -0,, ,.-


&8

7,0,0 77,., 7,7, ,5, ,,,,, 0,.7, ,.. 0 ..,, ,.,9, 07,0, ,..,

,
,..,, ,0.,, 9-5, ,...5 -, -. ..7, ,.,0 ,0,, ,.0.
=: to

0,0, 0,,5, .,., 07.,, ,. .,., ,.,0. 0,., ,.-97


<

5,,5, 8,,, 5,,5, ,,, 07,5, .,5, ,.-,, .,0, ,..,0 0,7, ,.,9.

,,, ,
, ,

,-,57
to

0,., 5,,5, ,,, 0,.7, ,99 ,,0, ,.-0. 0.,, ,.-,9

B0 7.95, 7,75, ,,.,, 88, -, ,,, ,75 ,,.5, ,.5,0 -.,,, ,.,,,
,, ,,, ,,

7,55, 7,,,, 7.0, .9.,, 7,,, 0..,, ,.,0, ,9, ..,, ,.5.5 .,,, ,.5,0

,, 0.,, 7,,5, ,,,,, 77.,, .7, ,.- -, ,0 .5. ,..55 -,,7, ,.5
,
,,, 7,, 0,, ,.0 -, -,, ..,, ,.0, ..,, ,.0.,

, ,,
,

,-5.
75, 7,8, .,7, 07,, ,.,. ,57 .,, ,.5,, 0. ,.,,,
264 THEORY OF FLOW AND FRACTURE OF SOLIDS

•fo
= 0.960 in the as-rolled and in the annealed condition, respectively (which figures
together with other observations show that the steel in the hot rolled condition was
already practically in an annealed state), while the steel of the former series of tests
broke in simple tension at a value -70 = 1.23. Although impact tests made on small
izod notched bars which were cut from the solid round bars parallel to the direction
of rolling and in transverse direction to it indicated that the impact strength was
40 per cent lower in the latter than in the former direction, the shearing stresses Tmw
at fracture given in Table 17-1 corresponding to an axial and to a circumferential
shear fracture in the hollow specimens did not differ much from each other. It is
believed that the scatter of the values of fmai at fracture obscured somewhat the
effect of anisotropy in the specimens due to rolling evidenced by the impact tests.

.t .3 .4 J
OCTAHEDRAL SHEARING STRAW

Fio. 17-33. Tests by Griffis and Morikawa on the yielding and fracture of 3-in. inside
diameter hollow cylinders of steel under biaxial stresses.

(As in the series of tests by Davis the tubes for a stress ratio n = \{at = <r«) still
broke along a crack in axial direction). It is interesting to note that in these hollow
cylinders of much larger size made of rolled stock compared with the steel tubes of
smaller size tested by Davis the shear fractures observed at room temperature for the
stress ratios n = 1,2, 00 soon after their formation changed to cleavage fractures. This
might have been caused through the less ductile behavior of the steel, but also through
the fact that with the increase of the geometric size of the specimen proportionally
more excess energy was available in the pressure fluid stored within the tubes of large
than in those of small diameter during the initial stages of the fracture process just
after the first leak developed through the shear crack. The plastic stress-strain
behavior of these tubes of 3 in. diameter agreed again very well with the common strain
hardening function r0 = f(fa) reproduced in Fig. 17-33, while a plot r«„ = /(fmu)
for these tests showed much more scatter of the observation points. A tube with
}<4

in. wall thickness developed cleavage fracture, while a tube with in. thickness
}$
a

and having the same inner diameter broke along a shear fracture although the true
stresses estimated at fracture were about the same in both tests.1 These and other
observations not reported here from this series of tests may throw some interesting

Obviously all these cleavage fractures developed after the steel had been deformed
1

permanently to a considerable degree during the distortion preceding the fracture.


TESTS UNDER COMBINED STRESS 265

side light on the conditions controlling the first formation and the propagation of the
fracture surfaces under biaxial stresses.
From the series of biaxial tests made at the University of California two investi
gations by Harmer E. Davis, G. E. Troxell, and E. R. Parker are quoted here. In the
first1 these authors reported about a group of preliminary tests made with thin-walled
seamless machined tubes of SAE 1010 and 1020 tested at room temperature and at
— 94°C under the stress ratios n = a, — 0,1,2.
/aa These specimens had 5.25 or
5.50 in. internal diameter, 0.160 in. wall thickness, and a gauge length of 11 or 22 or
25 in. and were tested in annealed condition. Small bars which were cut from the
tube wall and tested in tension gave practically the same true stress-natural-strain
curve as one of the longer hollow cylinders which was tested in simple axial tension.
Through the restricting influence of the short gauge length of 11 in. compared with

Table 17-2. Fracture Data in Biaxial Tests


(Harmer Davis, Troxell, and Parker)
Tubes 1020 steel; internal diameter 5.50 in., gauge length 25 in., thickness 0.16 in

Max elon
Test True stress gation, %
Stress ratio Wall thickness Mode of
temp., at fracture,
n — at/a. reduction, % fracture
°C lb/in.'
fa it

0 +21 111,000 102 —26 43 Shear


1 21 105,000 37 32 50 Shear
2 21 106,000 2 30 40 Shear
Various* 21 107,000 10 22 37 Shear
1 —94 63,000 2 3 5 Cleavage t
1 —94 74,000 2 2 3 Cleavage %
1 -94 80,000 5 5 10 Cleavage t
2 -94 138,000 0 50 37 Shear and cleav
age

* Stress ot maintained before fracture at ot — 53,000 lb/in. 2 and oa increased.

t Fracture occurred near weld near head.


X Fracture occurred in tube adjacent to weld.

22 in. length for the stress ratio n = 2 (internal pressure test) the maximum tangential
strain t, at fracture was reduced to 0.30 from 0.40 for the 22 in. long tube. Surpris
ingly the true fracture stresses at were practically the same for both tests. The true
normal stress (either aa or a,) for the 1020 steel tubes and the values of the axial and
circumferential (conventional) strains at the instant of fracture (measured in in.
x/i

gauge length) are reproduced in Table 17-2. With the exception of the short tubular
specimen in which the plastic deformation was restricted by the reinforced heads the
tests at 21°C exhibited a strain hardening behavior which correlated well with the
function t0 = /(-fo). The specimen tested with internal pressure only = at
2)
(n

— 94°C exhibited considerable plastic distortion before broke; the three specimens
it

tested under equal stresses at = aa{n = at — 94°C, however, disclosed little ductil
1
)

ity and broke in cleavage fractures which were influenced through the bending
stresses in the specimens near the reinforced heads (welds).

Progress Report "Tests on Small Tubular Specimens," October, 1944. N.D.R.C.


1

Project NRC-75 OEMsr-1221.


2GG THEORY OF FLOW AND FRACTURE OF SOLIDS

From explosions of large spherical industrial storage tanks of low-


carbon steel1 which occurred during recent years under service conditions
during the winter months, when the temperature of the surrounding air
was low, it is known that such steels (of a composition similar to the ship
plate steel) can fail in disastrous manner under static loading conditions.
During the past war years welded cargo ships suffered great damages on
the high seas when suddenly cracks appeared in the plates of their hulls,
sometimes breaking an entire ship in two halves.2 Most of these fail
ures in ships occurred at low atmospheric temperatures. The broken
pieces of the pressure vessels and the fractured surfaces of the plates of
the ships did not disclose the inclined shear fractures and contracted zones
characteristic of plates of this grade of steel stretched by tension or any
signs that the plates had appreciably permanently deformed near the
fractured surfaces as one would have expected to see for a ductile metal
and as one actually observes in the standard tensile acceptance tests
made on small round bars of these steels at these comparatively low
temperatures. On the contrary, the broken plates always showed the
characteristic "herringbone"3 pattern and cleavage fracture (see Sec.
15-6, page 191) normal to the plane of the plates.4

From the vast empirical information available in the literature on "notch sensitiv
ity" of steels, on the well-known drop of the impact energy in the so-called notched
bar test at temperatures near or below 0°C characteristic of the composition of regular
ship plate steels one would expect that, if small defects were present in the shells of
vessels or in the plates of ships, these might have contributed to the disastrous failures
at low temperatures in spite of the fact that impact forces were not present. The
welded seams between the plates are known to contain most of the defects in a welded
steel structure, and almost without exception all fractures of this type originated at
welds or near sharp fillets at reentrant corners (hatch openings in the decks of ships,

1 Brown, A. L., and J. B. Smith, Failure of Spherical Hydrogen Storage Tanks,


Mech. Eng., vol. 66, p. 392, 1944.
8 The United States Coast Guard has made available reports on these failures of

welded merchant ships.


3 Since the arrows in the herringbone pattern in broken steel plates point in the

direction from which the cleavage crack originated, it is easy to locate the point from
which the crack started.
* Having developed a method of producing the brittle fracture in ship plate steel

by the use of suitable notched test pieces having the thickness of the plate, C. F.
Tipper (The Fracture of Mild Steel, Proc. 1st Intern. Cong. Applied Mechanics, Lon
don, September, 1948) made valuable observations on propagation of brittle and
fibrous fractures, in which she could show that a number of crystal grains break by
cleavage and N'ewmann bands and that plastic deformation is confined only to a small
region on either side of the brittle fracture, also that the proportion of the two types
of failures (by cleavage and by elongation of a certain fraction of grains l>efore they
broke with a fibrous structure) varied with the degree of brittleness of the steel as a
whole.
TESTS UNDER COMBINED STRESS 267

along welded flanges in the pressure vessels). It is therefore certain that a combina
tion of the following factors was responsible for the disastrous formation of cleavage
cracks: (1) The presence of small holes or internal cracks in the welded seams or of a
concentration of stresses in reentrant corners or fillets. These alone may not endan
ger a construction of a ductile metal like plate steel under load if the temperature is
sufficiently high (20°C for ship plates) because the steel will locally deform by slip
within the zones in which the stresses are concentrated. (2) However, if the temper
ature drops below a critical value in the small previously strain-hardened zones
around the small holes or defects, conditions may become worse; a few crystallites of
the steel may find themselves under the condition in which they will break by cleav
age. In a ship hull riding the high sea waves great amounts of elastic bending energy
can be stored. Around the very sharp leading edges of fine cleavage cracks which
have formed at low temperatures the stresses under the sustained loads (large elastic
energies stored) will reach the dangerous values under which the cleavage cracks may
develop to herringbone fractures. (3) No "crack stoppers" were present in the
welded ships or storage tanks, in contrast to the riveted bodies of those which were
built in prewar times, which in their riveted joints contained them. (4) The types
of low-carbon steels used in these constructions are known to be "notch-sensitive"
(showing low impact energies below the critical temperature).

Leaving out entirely, for the present, further reference to these involved
questions, on which considerable work has been done,1 a second series of
tests on biaxial stresses may finally be mentioned here by Harmer E.
Davis, G. E. Troxell, and E. R. Parker,2 remarkable also for the full size
of %-in. ship plates of which the cylindrical specimens of 20 in. outside
diameter and 10 ft length were made. Twelve such cylinders were avail
able from the same grade of 0.23% C steel of which the 3 in. diameter
tubes tested by Griffis and Morikawa had been machined. The speci
mens were made from two half cylinders of 120 in. length welded along
two longitudinal seams 180 deg apart and along circumferential seams
joining the cylinder to reinforced hollow cylindrical end connections
(Figs. 17-34 and 17-35). An abbreviated summary of the test results is
reproduced in Table 17-3. From each of the ship plates of which the
large cylinders were formed small tensile test bars were machined (0.505
in. diameter, 2 in. gauge length) and tested at room temperature after
1 The brittle failure of steel plates containing welds or sharp artificial notches has

been the subject of many experimental investigations during the years 1943-1947.
Reference is made to work at the David Taylor Model Basin of the Navy Depart
ment in Washington, D.C., at the Universities of California (Berkeley) and of Illinois
(Urbana), sponsored by the National Defense Research Committee (War Metallurgy
Division) and of other institutions in the United States.
* Behavior of Steel under Conditions of Multiaxial Stresses; Tests of Large Tubular

Specimens," University of California, Berkeley; National Defense Research Commit


tee, War Metallurgy Committee, Final Report NRC-75, OEMsr-1221, August, 1945.
See also "Causes of Cleavage Fracture in Ship Plate. Flat Plate and Additional
Tests on Large Tubes," Report on Project SR-92, by Harmer E. Davis, G. E. Trox-
well, E. R. Parker, and A. Boodberg; August, 1946, p. 20, Appendix B.
268 THEORY OF FLOW AND FRACTURE OF SOLIDS

the same heat treatment had been applied to the large cylinders (except
the influence of welding). Although in these tensile tests with small
round bars a nominal tensile strength of 58,000 lb/in.2, a true average
fracture stress of 1 15,000 lb/in.2, an elongation of 40 per cent and a reduc
tion of area of 60 per cent were observed (these are approximate values

WELD

WIRE

Fio. 17-34. Longitudinal section through 20-in. steel cylinder.

neglecting to report here the differences of values which were recorded


between bars cut in the direction of rolling and in transverse direction to
it), the figures in Table 17-3 show that the welded large cylinders tested
under various conditions of internal pressure and axial load exhibited
much smaller values of the fracture stresses and of the reduction in
area.
The tubes tested at 70°F (with the exception of those which broke near
the ends because of disturbed stress conditions) nevertheless exhibited
sufficient ductility, but those tested at — 40°F behaved relatively in »
TESTS UNDER COMBINED STRESS 269

brittle manner (with the exception of the one which was heat treated).
At this low temperature the fractures were of the cleavage type and the
tubes shattered into fragments when tested at the stress ratio n — 1.
In all tubes in which the fracture initiated in the region of a weld, failure
occurred by cleavage, while shear fractures were observed first to form when
the failure did not originate in or
near a welded zone, but these shear
fractures frequently also developed METHYL ALCOHOL
into cleavage cracks. It is remark
PRESSURE
FLUID I

able that in the tests at 70°F in those -+-AIR CIRCULATED

cases in which the circumferential


stress was influential in causing the
fracture, the latter did not originate
in or near the longitudinal welded
seam in the cylinder, while in those
cases when the axial stresses were
TEST CYUMOCn
larger than the circumferential ones,
INSULATION
the failures initiated in or near this DRY ICE

seam. Thus the orientation of a AIR

major principal tensile stress relative SOLID CONCRETE


CYLINDCK
to the direction of a welded seam
determined whether the fracture orig
inated in the seam or elsewhere. A
Fia. 17-35. Longitudinal section
heat treatment given to the tubes through 20-in. steel cylinder.
after welding improved their strength
and ductility through alteration of the metallurgic structure rather than
through relieving of the residual stresses.
This series of careful biaxial tests on plate steel of normal thickness
shows that the data obtained in ordinary tensile tests on small specimens
overrate the strength properties of full-sized plates quite considerably,
and the strength values listed in Table 17-3 should be consulted for the
actual fracture stresses or strains in structures of the usual medium-
carbon steel. For many important details which influenced the forma
tion of the final fractures, which cannot be reproduced here, the reader
must be referred to the original reports. The cleavage fractures in several
of the broken cylinders disclosed the herringbone patterns which were
characteristic of the plates of ships that had failed in service.

Further tests under biaxial stresses on a magnesium alloy were made by D. M.


Cunningham, E. G. Thomsen, and J. E. Dorn1 using small tubes of 0.375 in. inside

' Proc. ASTM, preprint 1947. See also Thomsen, E. G., and J. E. Dorn, Effect
of Combined Stresses on the Ductility and Rupture Strength of Magnesium Alloy
Extrusions, J. Aeronaut. Sci., vol.
ii,

p. 125, 1944.
C
Table ,7,-. Biaxial Fracture Data for Large Welded Cylinder, of ,.0% Steel

(H. E. Davi,, G. E. T,ox,ll, a,d E. R. Pa,k,,, U,iv,,sity of Califo,,ia, B,,k,l,y, ,0,5)


Cyli,d,,,: OD ,, i,., thickn,s, in., gaug, l,gth ,, i,.*

Avg. true st,ss Max per cent


at fractu,, Max
elongation in Pot,
,duc

5
St,ss lb/in.> in. tial
Temp Heat tion in
Speci ratio, Loading End of t,atment wall energy Natu, of fractu,

n
men conditions connection test, thick at frac
(1,,°F)
"F Longi Trans Longi Trans t,e,
ness,
tudinal, verse, tudinal, verse, ft-lb

r.

u
at Co

A
Internal p,s With se0iellipti- , Befo, weld 420,, 00 <, ,5., 3,., ,33.,0J First shear, then cleavage
sure cal heads ing

2
II Internal pres With semicllipti- , After weld 45,,,0 .,,0 <, 2,.0 32., 1000,0 First shear, then cleavage
su, cal heads ing

B
2
S
\

Internal p,s With semiellipti- -02 Befo, weld 3,.,08 0,60 <0.2 3., 3., 1,.,,, Entirely cleavage
su, cal heads ing

C
||

i
Axial load, in Cylindrical head , Befo, weld 0,0, 8008 0.0 0.4 0., ,90,, First shear, then cleavage
ternal pressu, only ing

, ,
|1

D Axial load, in Cylindrical head , Befo, weld 0,0,; 66,,0! 3.8 0., 0.5 ,2,,0 Spirally running cleavage
'

ternal p,ssu, and ring ing

E
Axial load, in Cylindrical head, , Befo, weld 69,,,, 02,,,0 0.2 ,.6 2, ., 143.0, Cleavage around tube
ternal p,ssu, rinK, wi, ing

F
Axial load, in Cylindrical head, -40 Befo, weld 4600 4500, ,.6 2., 3.5 060,, Cleavage, tube shatte,d into
ternal pressu, ring, wi, ing 0any pieces
Axial load, in Cylindrical head, -30 After weld 85.50 88,0, ,0.4 ,8.0 3,., 2,,,,,0 Cleavage, tube shatte,d into
ternal pressu, rinK, wi, ing many pieces

I J
, , ,
Axial load, in Cylindrical head, -0 None 6,,,0 600,, 3.0 3.0 6.0 ,00,, Cleavage, tube shatte,d into
ternal p,ss,e ring, wi, many pieces

G
M
Axial load, in Head, ring -04 Befo, weld- 0,,0 25,,,, 2., ,.3 2., 0,,,0 Cleavage
ternal p,ss,e

0
Axial load, in -4, Befo, weld 43,16, 5,,60 4.2 3.8 6., 00,00 Cleavage fractu,
ternal p,ssu, ing

L
l l
Axial load, in -42 Befo, nnd 62,,,, 50,00 5.0 6.3 ,.8 2,0,0 Cleavage fractu,
ternal p,ssu, after weld
ing

*
is

Longitudinal section of specimen including end connection shown in Figs. ,0-30 and ,0-35.
Bold-face values of <raand at indicate direction of stress presu0ed to have caused fractu,.
Compression energy sto,d in pressu, fluid and in concrete plug (sec Fig, ,0-35) and elastic energy in tubular specimen,
Values of stresses a» and at are those fo, mid-section of tube at instant of fract,e.

j
Failu, occ,red through complex stress conditions caused through additional high

t J J|
at or near end connections of specimen and was presumably influ,ced
bending stresses induced by the ress ssnt of defoesation near heads. Values of m-n stresses at fract,e are 9gnificant, however, in that they indicate average
Rtrofuws responsible for initiation of fractu, in disturlxnl zones near hends.
TESTS UNDER COMBINED STRESS 271

and 0.437 in. outside diameter. They plotted T„t = /(Toot) curves for a number of
tests run under constant stress ratios n = <r,/at. Their observed points (root, Toot)
were found to he within a sufficiently narrow band around a common curve so that
this series of flow tests also agreed within the expected deviations arising from numer
ous causes with a postulated strain hardening function of the form just given.
In the series of tests under combined stresses described in the preceding pages, two
of the principal stresses were positive and of a tensile nature while the third principal
stress which was a compression stress was quite small and could in first approximation
be assumed to be zero. The points of the strain hardening function Toot = f(f<x>t)
which were experimentally determined in these tests were obtained under an essen
tially positive mean stress a = (<ri — <r2 + <r|)/3. The existence of the general strain
hardening function was based on the assumption not expressly mentioned that in
ductile metals Toot would not depend on the mean stress <r, whether positive or negative.
During the last war tests under varying conditions were carried out by P. W. Bridg-
man1 on ductile metals which may be characterized by stating that in them the flow
and fracture of metals (also other materials were tested by him) were investigated
under states of stress similar to some of the ones just mentioned after superposing
on these very high hydrostatic pressures. These tests therefore permit drawing
conclusions concerning the behavior of the investigated materials beyond the tensile
range of the principal stresses and permit a verification whether, for example, the
mean stress <r has an effect on the strain hardening function of ductile metals.
C. W. MacGregor and others have observed that the true stress & at a constant
rate of strain di/dt = const is a linear function of the natural strain in axial direction
(» = ci + c2i) in uniaxial tension tests beyond the load at which necking starts in
ductile metals until the fracture occurs. The flow stress a* is here defined as the differ
ence between the true axial stress aa and the radial stress e> at the distance r of a point
from the axis in the minimum cross section of the neck of a round bar. It can further
more be assumed in the plastic state that a = <r„ — a, = const (not dependent on r)
so that a also represents this difference of stresses at the axis (r = 0) of the bar under
which during the tensile test the fracture in the bar will start at the center of the
minimum section. The axial strain « and the two lateral strains according to Daviden-
kov2 can also be assumed to be constant throughout the entire minimum cross section
of the neck in the bar. The radial and tangential stresses <rr and <r, in the minimum
section of the neck change from the value zero at the circumference (r = a) to a
maximum value at the axis of the bar (r = 0) and must be equal to each other at
every point of this section. The maximum value of the axial stress aa at the center
r = 0 of the minimum section is the sum of the flow stress a and the maximum stress
a,. This stress <r„ causes the round bar to break at the axis r = 0.
Bridgman tested small cylindrical bars surrounded by a fluid in a container
in which very high pressures could be produced with suitable equipment. In
this manner Bridgman superimposed a constant high fluid pressure p on the small
specimens while they were stretched in tension. He found that the flow function

1 Eight reports submitted to the Watertown Arsenal, War Department, during the

years 1943-1944 and reports to the National Defense Research Committee, Div. 2.
Results of these investigations are published in the J. Applied Phys., vol. 17, pp. 201-
212, 225-242, 692-694, 1946; vol. 18, pp. 246-257. See also vol. 8, p. 328, 1937; vol. 12,
pp. 461-469, 1941; vol. 14, pp. 273-283, 1943. Reviews of Modern Physics, vol. 17,
pp. 3-14, 1945. Metals Tech., December, 1944, p. 32.
2 Loc. cit.
272 THEORY OF FLOW AND FRACTURE OF SOLIDS

after applying high fluid pressures around the tensile specimens remained practically
unchanged and points along the straight line 6* = ri + <"3« could be observed far
beyond the range of strains at which the bars would normally break in tension under
atmospheric pressure. From these tests it can be concluded that the strain hardening

4] 5
xl04kg/cme

Fio. 17-36.
Stresses at center of minimum cross section.

Fio. 17-37.
Stresses at surface of neck (r = a).
Flos. 17-30 and 17-37. The major principal stress circles representing the stresses •■"
fracture in Bridgman's tension tests for steel bars exposed simultaneously to high hydro
static pressure p.

function Toot = fifoct) does not change its shape appreciably under increasing high
mean pressures and that ■footdoes not depend on the mean stress a = (o-i — o-3 + «il/3
for the ductile metals which were investigated.
When the specimens were pulled in the pressure chamber under the superposed
high fluid pressure p they necked down much more than when p = 0 because the
high lateral compression stresses reduced the axial tensile stresses around the circum-
TESTS UNDER COMBINED STRESS 273

ference of the minimum section in the neck required for a plastic distortion. Natural
strains in the axial direction were observed of the order of la = 2 to t„ = 4 (corre
sponding to conventional strains « = 6.4 to 54). In the fracture surface of the
specimens the outer conical portion increased and the inner portion having a fibrous
structure decreased with the increasing pressures p. Above 15,000 kg/cm* pressure
the inner fibrous part of the fracture surface no longer appeared and the entire frac
ture consisted of one cone (or of two cones) or of a small plane area, inclined at approxi
mately 45 deg with respect to the axis of the bar. While the tensile axial stress aa
at the bottom of the neck (r = a) decreased, the axial stress a„ at the center of the
minimum cross section (r = 0) increased to high positive values with increasing
pressures p at the instant of fracture. The maximum shearing stresses
Tmm, = »/2 = (<r„ — o>)/2

at fracture may be represented by the radii of the major principal stress circles. From
the data of these tests1, two figures have been constructed (Figs. 17-36 and 17-37)
indicating the major principal stress circles and their relative positions in the o,t
Mohr stress plane at the instant at which the round bars of one of the investigated
steels had broken. The corresponding fluid pressures p which were reached are
inscribed below the stress circles. Figure 17-36 represents the states of stress
at the center r = 0 of the minimum cross section in the neck and Fig. 17-37
illustrates the Mohr circles of a material point at the circumference r = a of
the neck at the instant of fracture. Because of the excessive cold work that had
been concentrated within a comparatively very small portion of metal near the
minimum cross section in the neck of the bars and considering the very steep gradients
of the stresses <ro and a, in the radial and axial directions and the rapid changes of
the permanent strains i in the axial direction (in those bars that necked down nearly
to a point at the highest pressures p) it appears of doubtful value to delve into special
criteria of fracture based on stress or strain that might be exhibited in comparing a
few quantities deduced from these tests. Reference in this respect may be made to
previous remarks in Sees. 15-5, 15-6, and 15-9 in which the complexity of the factors
influencing the conditions of fracture was mentioned.2
1 Taken from the tests of one of the investigated steels (Fig. 15, p. 210, of the paper
published in the J. Applied Phys., vol. 17, 1947).
* In some of the earlier reports which were quoted on p. 271 Bridgman stated that

a constant value of the mean stress (<rt + <rt + <rt)/3 would determine the condition
of fracture at the center point of the minimum cross section of the bars which were
broken in tension and under high lateral pressures. In a paper published in 1946
(J. Applied Phys., vol. 17, p. 212), however, he writes, "A search is made for a possi
ble criterion of rupture by plotting various relevant stresses against the strain at
fracture. No criterion emerges good under all conditions. The criterion of the mean
hydrostatic tension (one-third the sum of the three principal stress components)
remains the best criterion over the whole range of conditions, but in certain circum
stances it may show considerable variation, and its margin of superiority over the
criterion that the total fiber stress in the direction of fracture be constant at fracture
is not impressive." The criterion of a constant value of the mean stress fails when
states of stress are compared in which the two minor Mohr stress circles have equal
radii, i.e., when the intermediate principal stress a% is the arithmetic mean of ai and
o•2. Under a state of pure shear oi = t, <r2 — 0, <rj = — t the metal breaks at a cer
tain value of r, but at a vanishing value of the mean stress.
For further remarks concerning Bridgman's assertion that a constant limiting
274 THEORY OF FLOW AND FRACTURE OF SOLIDS

The conditions under which metals were deformed in these severely cold worked
zones might be crudely compared with those encountered when a piece of a ductile
metal is rolled between rigid rolls or drawn through a set of cylindrical dies in repeated
operations without being annealed and when it is simultaneously pulled in the direc
tion of positive strain by a tension force. Practical engineers make use of combined
tension and lateral compression for forming metals hot or cold and have noted that
the very heavy lateral rolling pressures required to produce large cold reductions in
repeated single passes against the energy-consuming large friction forces resisting the
flow may be materially reduced when a strip passing between the rolls is simultaneously
pulled in the direction of rolling or by a back tension or pulled in both directions.
Conversely, the tensile stresses stretching a strip at its necking point are reduced
when it is simultaneously compressed between revolving rolls in lateral direction.
The condition in which the grain structure of a ductile metal is left after a series of
such operations is not sufficiently known. In practice, a certain maximum amount
of accumulated cold work may be desirable or permissible; beyond that amount strips
or wires must be annealed, not only for stress relief, but because of the expected
deterioration of the tensile properties of the metal through too much cold work.

value of the mean stress (<n + a2 + ai)/Z would determine the criterion of fracture,
reference is made to papers by D. J. McAdam and his associates (particularly to
McAdam, G. W. Geil, and F. J. Cromwell, Flow, Fracture and Ductility of Metals,
Metals Technol., No. 2296, p. 24, January, 1948), who also reject this criterion. The
latter paper contains valuable test results on the influence of previous cold deforma
tion of metals on the fracture under combined stresses.
CHAPTER 18

STRAIN AND FLOW FIGURES

Before I began the study of electricity I resolved to read no mathe


matics on the subject till I had first read through Faraday's Experi
mental Researches on Electricity.
For instance, Faraday, in his mind's eye, saw lines of force traversing
all space where the mathematicians saw centres of force attracting at
a distance: Faraday saw a medium where they saw nothing but dis
tance: Faraday sought the seat of the phenomena in real actions
going on in the medium, they were satisfied that they had found it in
a power of action at a distance impressed on the electric fluids.

Maxwell, J. Clark, Preface, "Treatise on Elec


tricity and Magnetism," Vol. 5, 3d ed., 1892.

Strain Figures in Mild Steel.


18-1. If a tensile test piece of mild steel,
especially a flat bar, is given a mirrorlike polish and then tested in tension
in a testing machine, at the instant of the drop in the load at the yield
point often fine dull lines appear on the polished surface of the bar at an
angle to the axis of tension. These lines, known to engineers as flow or
strain figures or as "Luders' lines," quickly spread over the length of the
bar while at the same time their thickness increases. The observation
of the occurrence and spread of these fine markings on the surface of steel
test pieces gives valuable information on the phenomena taking place in
the structure of the material at the instant it yields.1
On account of the regular occurrence of these flow figures on stressed
specimens of mild steel and of the remarkably regular orientation of these
layers of slip with respect to the directions of the principal stresses it is
of a greater interest to study how they originate, because of the close
relation which seems to exist between the orientation of these thin layers
of slip and the state of stress in a steel piece stressed to the plastic limit.
1 The
flow figures seen in soft steel were first described by Luders, Dinglers Polytech.
J., " Materialienkunde,"
1854. Cf. also Martens-Heyint Berlin. The French
artillery officer, L. Hartmann, appears to have been the first to study these lines
thoroughly in his book, "Distribution des Deformations dans les M£taux Soumis a
des Efforts," Berger-Levrault, Paris, 1896. The importance of these lines for the
mechanics of the plastic state of metals was recognized by Otto Mohr, Z. Ver. deut.
Ing., 1900. The formation of the flow figures was studied by several investigators,
among whom T. H. Turner and I. D. Jevons, J. Ill,
Iron Steel Inst., vol. No. 1,
p. 169, 1925 and E. W. Fell, ibid., 1927, may be mentioned here.
275
276 THEORY OF FLOW AND FRACTURE OF SOLIDS

On bars of mild steel covered with a coating of rust or mill scale, the flow
or strain figures may also be observed. If such bars have been deformed
beyond the yield stress the scale begins to flake off and thus regular
markings on the surface of bars may be seen, which are identical with
these flow figures. Some examples of flow
figures on polished specimens are shown
in Figs. 18-1 to 18-3.

In a tensile test the first flow lines generally


appear suddenly, usually at or near points on the
piece where it begins to enlarge near the heads of
the piece. An example of the way these lines
appear is shown by the sketches of both flat
sides of a test piece in Fig. 18-4. The first line
to occur was that represented by the black line
o. At the instant that this line became visible,
the load on the test piece which was at the
"upper" yield point dropped about 5 per cent
Fio. 18-1. Strain or flow figures and, while the load dropped 2 per cent more, a
(Liiders' lines) on compressed steel
specimen. second line b formed. From these two lines there
gradually spread out, under constant tension, a
wide dull band c (see Fig. 18-4). Two such bands are often observed at the yield
point in positions as shown in Fig 18-5, the middle portion of the test piece having
shifted sideways a small amount, In the case of compression tests with mild-steel
specimens a similar phenomenon may be observed as shown in Fig. 18-6.

Fio. 18-2. Fig. 18-3.


Figs. 18-2 and 18-3. Strain or flow figures (Liiders' lines) on compressed steel specimen.

In the narrow dull strips or lines, which mark the position where the plastic layers
intersect the surface of the test piece, the direction of the relative movement of the
unchanged bounding portions of the material, with respect to each other or the slip
layer, may be recognized. These lines or strips may have either the profile of a
shallow groove (Fig. 18-7o) in case of tension, or of a shallow ridge in case of com
pression (Fig. 18-7c), or they may form a flat slope (Fig. 18-76), which appears scaly
under the microscope. The grooves appear where the dull strips run in a direction
about 45 deg to the axis of the test piece, while the flat slopes form where the dull
STRAIN AND FLOW FIGURES

strips are perpendicular to the direction of tension. Where the direction of slip runs
parallel to the surface of the test piece, these shallow grooves result. In a steel test
piece under compression, similar Mow lines likewise result. In such cases, however,
the flow lines have, instead of the profile of shallow grooves, that of a flat ridge

4
v \V
Fig. Fig. Fig. 18-7.
18-5. 18-6.
Fig. 18-4. Strain Figs. 18-5, 18-6, and
figures on a flat 18-7. Formation of
bar of steel. slip layers by tension
or compression.

Klg. 18-8. Micrograph of the slightly distorted surface of a well-polished specimen of


mild steel just after the yield stress has been reached. Note the tendency of the slip bands
appearing in the crystal grains to orient themselves parallel to a certain direction. The
micrograph was taken along the border line of a strain figure (Luders' line).

(Fig. 18-7c). Under the microscope the flakes or scales visible in these flow lines
often prove to be groups of crystal grains displaced along neighboring layers and
deformed plastically. In the grains dark wavy lines may be recognized (see the
photographs of Figs. 18-8 and 18-9). These markings may be identical with the
wavy lines which Taylor and Elam1 found in plastically deformed single crystals
of iron.
1 Proc. Roy. Soc. London, ser. A, vol. 112, p. 337, 1926.
278 THEORY OF FLOW AND FRACTURE OF SOLIDS

The production of the shallow grooves on the surface of a stretched test piece of
mild steel and the flat ridges on the surface of a steel prism loaded in compression is
apparently the consequence of local yielding. The local axial strains equal about
2 to 4 per cent and the load after the sudden drop at the yield point remains
nearly constant while the test specimen stretches. A permanent extension of 2 to
4 per cent in the axial direction in a tensile test corresponds to a lateral contraction
of 1 to 2 per cent. An oblique section of the test piece taken at an angle of 45 deg
to the direction of tension, the trace of this section on the surface being the groove of

Fia. 18-9. Micrograph of the structure of a steel bar, bent permanently to a small degree.
The micrograph was taken along the border line of a strain figure (Liiders' line). Observe
that the crystallites slipped together in groups and were displaced in parallel layers. In the
deformed grains slip bauds appear. (Magnified about 70 times.)

the flow line, must therefore contract about 1 to 2 per cent while the bounding part
of the test piece does not change its dimensions.

One property of the flow or slip layers has an especial bearing on the
mechanics of the plastic state. The planes of the thin slip layers in
which the iron is apparently more severely deformed than elsewhere
coincide approximately with two of the planes of principal shearing stress.
The angle a which the slip layers make with the axis of the test piece is
in the case of tension usually a little greater (47 deg) and in the case of
compression usually a little less than 45 deg1 (see Figs. 18-10 and 18-11).
Since these properties of the flow layers also are exhibited in complicated

cf. Scholl, Versuche iiber Gleit- und Brucher-


1 For measurements of this angle

scheinungen, Z. Ver. dexii. Ing., p. 406, 1925.

-
STRAIN AND FLOW FIGURES 279

three-dimensional distributions of stress, observations relative to their


position form a valuable aid in the investigation of states of stress in
plastically deformed solids which will often be used in what follows.
The brittle layer of scale, which covers hot rolled or forged steel bars,
flakes off frequently where these bars have subsequently been stressed
in the cold state above the plastic limit. In the neighborhood of punched

Fia. 18-10. Fio. 18-11.


Figs. 18-10 and 18-11. Orientation of the slip planes in a tensile and a compression test
piece.

rivet holes, markings of considerable regularity are often revealed These


lines are identical with Luders' lines.
For observing the flow figures, three methods are available. In the
first method, which was worked out by A. Fry,1 for steel the test piece
after being stressed is heated to 200 to 250°C, after which it is etched by a
very strong solution containing hydrochloric acid and copper chloride.
The traces of the flow layers on the cross section appear as dark lines.
Fry and Strauss2 have been able to obtain, with the help of this etching
method, valuable information regarding the fundamental changes which
iron undergoes at the plastic limit.
1 Kraftwirkungsfiguren in Flusseisen, dargestellt durch ein neues Xtzverfahren,
Krupp. Monatsh., July, 1921; also Stahl u. Eisen, 1921.
1 Krupp. Monatsh., July, 1921.
280 THEORY OF FLOW AND FRACTURE OF SOLIDS

A second method to make the flow lines visible is to photograph the


flat relief shown by the depressions and ridges on a finely polished
metal surface by means of the Topler "Schlierenmethode," which was
applied by the author at the suggestion of L. Prandtl.1 The photographs
of slightly distorted metal surfaces, given frequently throughout this
book, were taken by the use of this method.
The third method utilizes the observation which has been mentioned
that the brittle iron oxide scale covering hot rolled pieces of steel flakes off
in those portions of the surface below which the steel has locally yielded.
At the yield point of a bar or prism of soft steel the strains in the thin flow
layers in a tension or compression test increase from an elastic strain of

Fig. 18-12. "Schlierenmethode" for observing flow or strain figures on polished metal
test pieces. Li, source of light. S, mirror. Pi, prism. Li, large lens. P2, test speci
men. K, camera. M, ground glass. B\ and Bj, diaphragms.

the order of one-thousandth to values of the order of 0.02 to 0.04. The


iron oxide scale cannot sustain such large strains and flakes off at the
intersections of the flow layers (Liiders' lines) with the surface. By paint
ing the scale with a whitewash the flow layers can be made sharply visible
as dark bands. Recently a special grade of the "stress coats" (used for
measuring surface stresses in the elastic range of loaded pieces) has been
developed by the Magnaflux Corporation in Chicago, 111., with which the
machined surfaces of specimens can be painted and which can serve to
bring out fine details of the slip lines in mild steel.
The etching method of Fry is particularly suitable for studying the
distribution of the surfaces of slip in the interior of mild steel specimens.
The two other methods are conveniently used to make them visible on the

1 Schweiz. Bauztg., vol. 83, Nos. 14 and The author is indebted to Dr.
1924.
15,
Lihotzky of Wetzlar for optical equipment which this purpose very well
has served
and which was attached to a metallic microscope manufactured by Ernest LeitE of
Wetzlar; this arrangement is represented in the sketch of Fig. 18-12. In this sketch,
Li is an arc lamp or source of light, S a mirror, Pi a totally reflecting prism, L. a
good lens with a large aperture, P2 the test specimen, K a photographic camera, M a
ground glass, Pm and B% two diaphragms. In operation the screen opening B\ illumi
nated by the arc lamp is projected by means of the lens Li and the reflecting test piece
Pi on the plane of the screen B2.
STRAIN AND FLOW FIGURES 281

Fig. 18-13. Fig. 18-14.


Figs. 18-13and 18-14. Meteoric iron. Compression test (note Widmannstiidten struc
ture in right view). (Cotirtesy of Ninninger Laboratories, Denver, Colo., 1932.)

Fig. 18-15. Fig. 18-10.


Figs. 18-15 and 18-16. Luders' line pattern on a steel sheet. After the thin steel plate
was stretched by a small strain (in the direction parallel to the short sides of the photo
graphs) the scale fell off, disclosing the flow lines. The left and right pictures are views
of the front and rear side of the steel sheet. Note the perfect correspondence of the two
patterns of flow lines indicating that the slip planes were perpendicular to the plane of the
sheet. (Acknowledgments are due to F. C. Biggert, Jr., and Dr. M. Stone of United Engineer
ing & Foundry Company, Pittsburgh, Pa.)

surface of stressed pieces. Examples of the flow layers developed by the


Fry method are given in Chaps. 20 and 36. The two photographs repro
duced in Figs. 18-13 and 18-14 illustrate the application of the Schlieren
method. Figures 18-13 and 18-14 show the two opposite faces of a small
prism of square cross section of a very soft meteoric iron that has been
stressed in compression. The two faces were polished to a mirror finish
282 THEORY OF FLOW AND FRACTURE OF SOLIDS

before the compression was applied. The prism contained a small hole
in the center in perpendicular direction to the direction of compression.
The photographs show the finest disturbances in the mirrorlike surface
greatly exaggerated including polishing scratches. Figures 18-15 to 18-18
illustrate the pattern of Liiders' lines on a thin sheet of steel covered by
the iron oxide scale that has been pulled in a steel mill cold by a small
amount, on a flat bar of silicon iron (the scale on it from the hot rolling

Fio. 18-17. Liiders' lines on a flat bar Fig. 18-18. Thin plate of silicon iron bent
of silicon-steel tested in tension. under concentrated load in center. The
edges were simply supported.

was painted with a white color) after the yield point has been reached in
tension and on a small rectangular plate of the same steel which was
loaded by a concentrated force in the center. This rectangular plate was
freely supported along the four edges, but after the plate started to yield
the corner regions were lifted up from the frame on which they rested.
It may be noted that the rectangular plate was essentially plastically
strained in bending in a zone bridging the center regions of the two longer
edges of the rectangle. Figure 18-19 reproduces three plastic "wedges"
(groups of families of Liiders' layers) which have formed on a flat steel
STRAIN AND FLOW FIGURES 283

bar pulled eccentrically by equal tension forces in the direction parallel


to the axis of the bar. In this case the fine markings were made visible
by a softer grade of Magnaflux stress coat.1

Fio. 18-19. Wedge-shaped plastic regions in a flat bar of steel (width 1.5 in., thickness
0.125 in., prismatic gauge length 5.5 in.) which was subjected to an eccentrically applied
tensile load. The load acted in a line parallel to the axis of the bar in a distance (eccen
tricity) of 0.250 in. to the right of the axis. The fine slip lines were brought out by using a
stress coat. J.
(.Testa by Miklowitz.)

In the thin flat bars tested in tension similar to the one reproduced in
Fig. 18-17 these very sharply defined thin parallel flow layers developed
successively from one or both ends of the bar and were perpendicular to
1 The
formation of these wedge-shaped plastic zones consisting of numerous thin
flow layers was described by J. Miklowiz in his paper, The Initiation and Propaga
tion of the Plastic Zone in a Tension Bar of Mild Steel under Eccentric Loading, J.
Applied Mechanics (ASME meeting, December, 1946).
284 THEORY OF FLOW AND FRACTURE OF SOLIDS

the flat sides of the tension bars. Their traces appeared on both flat
sides, and in each flow layer the plastic strain was a simple shear (a plane
strain).1 It is remarkable that in the case of the flat silicon steel speci
mens the parallel slip lines developed at certain equal intervals, as can
be seen in Fig. 18-17. The dark thin flow lines did not become thicker
after they appeared on the sides of the flat bar, but new finer lines
appeared between them later. The thickness of the flow layers that
form in flat bars or sheets of steel seems to be proportional to the thick
ness of the specimen.
In certain steels, however, a gradual dulling of the polished surfaces
is often observed at the yield point, and the boundary of the dull (yielded)
zone in flat bars is seen to advance as an oblique line with a certain veloc
ity which depends on the speed of the moving head of the testing machine.
Much is yet unknown concerning the mechanism of the formation of the
flow layers in tension tests of mild steel. So much is established that a
number of important mechanical factors influence their mode of forma
tion : shape of specimen, of connection with heads, the degrees of freedom
available for the motion of the heads or grips of the testing machine, the
eccentricity of the load ; the stiffness of the testing machine (expressed as
AP: Ax, P load, x relative displacement between heads of machine, assum
ing that the specimen is perfectly rigid), grain size of steel, and conditions
in which the perlite is distributed in or around the grains (a fine-grained
normalized mild steel shows, usually sharply, a coarse-grained steel hav
ing a gradual transition curve from the elastic to the plastic portion in the
stress-strain curve dull flow layers). Investigators have frequently been
puzzled by the observation that the Luders' lines which were visible on
polished specimens could not be developed by the etching method of Fry
in specimens of the same steel. This was explained by W. Koester after
he showed that the etching method responds particularly well on steels
containing small amounts of dissolved nitrogen.2

1 The gradual spreading of the Liidcrs' lines was studied in motion pictures. A
film on their formation was shown at the International Congress for Testing of
Materials in London in 1937.
* A low-carbon
steel, according to W. Koester (Zur Frage des Stickstoffs im tech-
nischen Eisen, Arch. EisenhiUtenw., vol. 3, No. 10, p. 637, April, 1930) can dissolve
0.02 per cent nitrogen at 400°C and only 0.001 per cent at 100°C. Under normal
conditions steel at room temperature represents a supersaturated solid solution of
nitrogen. (It requires long times of heating at 100°C to eliminate the nitrogen from
steel.) The Fry reagent reacts particularly well with nitrides that have been precipi
tated out in the deformed zones in the flow layers. In the undeformed zones, this has
not been the case after the short heating of % hr at 200°C required for the general
developing of the Li'iders' layers. Artificially nitrated steels blacken by the Fry
STRAIN AND FLOW FIGURES 285

According to tests by P. Ludwik and R. Scheu1 the yield point of


electrolytic iron may be raised by combined mechanical and thermal pre
liminary treatment. They emphasized that there are other factors than
the carbon content which may greatly affect the properties of steel,
causing a high ("upper") yield point. Among these attention was called
to the minute quantities of dissolved gases which seem to affect the form
of the transition curve to the plastic range of strains and to be the causes
of the phenomena known as the "aging" of steel.2
A. Fry3 distinguishes three kinds of disturbance in the microstructure
of mild steel that has been deformed in the flow layers at the yield point :
disturbances along the grain boundaries (the latter appear much thicker
and darker in the micrographs etched by means of his reagent within the
flow layers than in the elastically deformed parts of the steel pieces) ; dis
integration of the grains (cracks and slip lines are visible within the
individual ferrite grains) ; and, finally, certain disturbances in the grains
themselves, distributed uniformly over their sections appearing in the
micrographs. (The plastically strained ferrite crystallites within the flow
layers are covered by fine dark spots and appear much darker etched than
the ferrite grains in the adjoining elastically stressed portions of a speci
men). These observations support views which were expressed on page
45 that in the flow layers of a mild steel the skeleton of the intercrystal-
line substance (the cementite or pearlite) breaks first when the yield point
is reached.4 The collapse of this strong skeleton is the main reason why a
sharp yield point in low-carbon steel is observed. After the resistance

reagent; recrystallized grain and certain nitrogen-poor steels on the contrary do not
react. This then explains the observations which were quoted above and the neces
sity of heating previously deformed specimens containing sufficient nitrogen to
200°C for hr before the Luders' lines can successfully be etched. See also the paper
by C. W. MacGbeoor and F. R. Hensel,, The Influence of Nitrogen in Mild Steel
on the Ability of Developing Flow Layers, /. Rheol., vol. 3, No. 1, pp. 37-52, 1932.
In their test they found that the Luders' lines are best brought out by the Fry reagent
on Bessemer steel containing a certain favorable amount of dissolved nitrogen, while
open-hearth steels reacted less favorably. Artificial nitrating enhanced the etch-
ability in steels.
1 tlber
die Streckgrenze von Elektrolyt und Flusseisen, Werkstojfausschussber.
No. 70, 1925; also Stahl u. Eisen, 1925; Z. Ver. deut. Ing., vol. 70, p. 379, 1926.
*Cf. aging, see Fettweis, F., Stahl u. Eisen, vol. 39, p. 1, 1919; vol. 42, p. 744,
1922. Also Koerber, F. and A. Dreyer, Stahl u. Eisen, vol. 43, 1923 (on the blue
brittleness in steel).
5 hoc.
cit.
4
H. Hanemann, A. Schrader, and W. Tangerding (Arch. Eisenhuttenw., vol.6,
p. 567, 1932) could detect under large magnifications a fine layer around the ferrite
crystallites which appeared broken.
286 THEORY OF FLOW AND FRACTURE OF SOLIDS

Fig. 18-20. Fig. 18-21.


Figs. 18-20 and 18-21. Compression tests with paraffin prisms containing a cylindrical
hole at right angle to direction of compression. Left: beginning of distortion near hole.
Right: formation of two distinct slip layers starting from hole.

Fig. 18-22. Fig. 18-23. Fig. 18-24. Fig. 18-25. Slip


Figs. 18-22, 18-23, and 18-24. Slip lines on the sides of a layers in cylinder
paraffin prism having a hole, and stressed in compression. with hole.

of the reinforcing frame around the ferrite grains has been diminished, the
latter are severely deformed by shear and their lattice structure is further
damaged within the regions of the thin layers in which the plastic distor
tion is localized.1
1 A. Eichinger (Arch. EisenhuUenw., vol. 18, pp. 73-90, 1944), in contrast to the
views which were just expressed, attributes the sharply defined yield point of mild
steel to the strengthening effect of heterogeneous atoms which have been precipitated
out in the a iron along the faces of the small blocks constituting the "mosaic struc
ture" of the single grains of the ferrite. The impurities assembled along the mosaic
blocks within the crystallites would increase the resistance of the atom layers against
slip. X-ray reflections observed by F. Lihl (Metallwirtschaft, Nos. 23, 24, pp. 391,
49, December, 1948) seemed to support the hypothetic presence of a reinforcing
frame or skeleton of atoms within the faces of the secondary mosaic structure of the
ferrite grains of a mild steel. According to J. H. Palm (Natl. Aeronatdical Research
Inst., vol. 15, 1949, Rept. M. 1230, Amsterdam; see also Metalen, vol. 3, No. 5,
STRAIN AND FLOW FIGURES 287

Valuable observations about the formation of the flow layers in steel


and in other metals were described by T. H. Turner and J. D. Jevons1
and by E. W. Fell2 in their papers, to which further reference is made.

18-2. The Production of Flow Lines by Notches and Holes. As is known from the
theory of elasticity, the stresses in elastic bodies are increased or concentrated at con
cave boundary surfaces, for example, at the edges of a cylindrical hole or a spherical

Flo. 18-26. Two views of the same paraffin cylinder having a small hole. The cylinder
has been subjected to compression in axial direction. The slip layers in the left view form a
white cross and appear as ellipses in the second view. (The bright central line parallel to
the axis of the cylinder is caused by a reflex of light and not by deformation.)

cavity in a member subjected to pure tension, compression, or shear. At a sharp


reentrant edge the stresses are theoretically infinite. Even if the stresses in the
neighborhood of a reentrant edge have only small values, nevertheless, these may be
sufficient to produce plastic deformation or even fracture at the corners.
In compressed test pieces of brittle material, such as cast iron or hard steel, which
contain small notches, failure by fracture begins at these notches.3 By means of
small cylindrical holes in an area subjected to pure axial tension or compression, cer-

January, 1949) the dissolved carbon and nitrogen atoms which do not occupy posi
tions of the atoms in the lattice of the ferrite crystallites by "substitution" but by
"interposition" between them are the cause of the yield-point phenomenon. See also
Low, J. R., and M. Gensamer (Metals Techno!., Tech. Paper No. 1644, December,
1943) on the related phenomena of the strain aging of steels.
1 The Detection of Strain in Mild Steels, /. Iron Steel Inst., vol. Ill, No. 1, pp. 169-

213, 1925.
! The Piobert Effect in Iron and Soft Steel
(1935) and Yielding Phenomena and
Distortion in Iron, Steel, Aluminum Alloy and Other Metals under Stress (1937), J.
Iron Steel Inst.
2 Leblond, Tech. moderne, vol. 15, p. 7, 1923.
288 THEORY OF FLOW AND FRACTURE OF SOLIDS

tain disturbances may be produced in plastic materials, as shown in Figs. 18-20 to


18-32.
The local yielding in the neighborhood of a hole or a
notch may be studied by means of a test piece made of a soft
material.1 For this purpose test pieces made of soft paraffin
and having holes were stressed in compression, the compres
sion being perpendicular to the axis of the hole. Long before
the first slip lines were visible the slight deformation of the
surface of the test piece showed that the plastic deformation
was not uniform but occurred mainly in two planes. These
two planes intersected along the axis of the hole and formed
equal angles with the direction of the compression. The first
distortion of the surfaces of a compressed prism of paraffin is
shown in the photograph of Fig. 18-20. After an increase in
loading there resulted the flow lines shown in the photograph
of Fig. 18-21 and indicated in the sketches of Figs. 18-22 to
18-24. The test piece split, so to say, into four rigid pieces,
separated by two soft plastic layers, as indicated in Fig. 18-24,
the wedges o and 6 moved in a vertical direction, thus forcing
the wedges c and d to separate in a horizontal direction.
Similar phenomena in a paraffin cylinder, having a hole per
pendicular to the direction of compression, are illustrated in
both views of Fig. 18-26 and the sketch of Fig. 18-25. The
Fia. 18-27. Distor
tion of side of a flat slip planes appear as white strips in the photographs and are
steel bar having two likewise grouped about two planes as indicated. The effects
very small grooves on of a small notch or a small hole in producing local yielding
edge. Tensile test.
in two thin layers in metal specimens under compression or

Fio. 18-28. Fio. 18-29. Fig. 18-30.


Figs. 18-28, 18-29, and 18-30. Slip layers produced by two small grooves (Figs. 18-28 and
18-29) or by central hole (Fig. 18-30) in hard copper prisms loaded parallel to edge in
compression.

tension are shown in the photographs Fig. 18-27 (steel), Figs. 18-28, 18-29, and 18-30
(copper), and for paraffin in Figs. 18-31 and 18-32.
1 S. Timoshenko utilized Liiders' lines to indicate the instant at which a definite
maximum stress at a dangerous point in a stressed body is exceeded; cf. Proc. 2d
Intern. Congr. Applied Mechanics, Zurich, 1926.
STRAIN AND FLOW FIGURES 28'.)

18-3. Elastic Stress Distribution and the Beginning of Plastic Flow


in a Plate with a Circular Hole.1 In machine construction, tension or

Fig. 18-31. Fig. 18-32.


Plus. 18-31 and Paraffin.
18-32. Formation of slip layers crossing at approxinately
right angles. Produced by an accidental bubble (left); by a small hole (right).

compression members having holes are very frequently used. It is


important to note that the effect of a small cylindrical hole in such mem
bers is quite different, according to whether
the material around the hole is stressed only
elastically or whether it is partially yielding,
plastic deformation occurring under a sharply
determined yield stress.

The stress distribution in a tension or compres


sion member in the form of a wide plate containing
a small cylindrical hole is well known in case of an
elastic material.2 Referring to Fig. 18-33 consider
a point P, located in the thin plate having a hole of
radius a. The polar coordinates of this point are r
and <j>. If the plate is subjected to a uniform tensile
<f>

stress s in the direction of the polar axis = the


0,

three components of stress which determine the state


of stress in the plate are given by the following Plate with hole in
expressions:

"'j-i'i 4a» 3a<\ „


1

Radial stress: a.

Tangential stress: at (18-1)

Shearing stress: xr(

See Baud, R. V., Wahl, A. M., and author. Mech. L'ng., p. 187, March, 1930.
1

This problem was first solved by B. Kirsch, Z. Ver. deut. Ing., July 16, 1898,
*

p. 797. Further applications of the theory of elasticity to the case of an elliptical


hole were made by Inglis and Wolff.
290 THEORY OF FLOW AND FRACTURE OF SOLIDS

On the circumference of the hole the coordinate r is r = a and hence the stresses
become equal to
a, = o , <r, - «(1
- 2 cos 20) , t>, = 0 . (18-2)

The radial and shearing stresses a, and r,< vanish. The tangential normal stress a,

<t>
= = 3»72, located on the circumference of the
is a maximum at the points x/2 and

Kig. 18-34. Elastic plate with hole in tension. Contour lines indicate curves as computed
according to Eqs. (18-1) and (18-5) along which maximum shearing stress Tm»i is constant.
These curves correspond in a photoelastic test to the "isochromatic lines."

hole and on an axis of the plate perpendicular to the direction of the tension. At these

points the tangential normal stress


is

<Trau = 3» . (18-3)
<f>

For = = = *, however, at becomes negative


and or (i.e., a compression)


a

0
r

and
a, = -s. (18-1)

These formulas indicate that the presence of small hole in an infinite plat* of an
a

elastic material subjected to a pure tension stress in given direction causes con
a
a

siderable increase in the stresses in the vicinity of the hole, the maximum normal stress
on the boundary of the hole being three times as large as the stress in an undisturbed
portion of the plate.
To confirm by test these analytical results and to describe the distribution of stress
the curves can be calculated along which the maximum shearing stress rm has cer
a
STRAIN AND FLOW FIGURES 291

tain constant value.1 The equation of these curves is given by

Wi
- Or)' = k' -T- = const
+ Trl* (18-5)
4

where k is a constant. Using here the expressions (18-1) the contour lines of the dis
tribution of the maximum shearing stress ru, as a function of the polar coordinates r
<t>

and can be determined. These lines are plotted in Fig. 18-34. For example, the
contour line marked = the locus of the points where the ratio of T,n„ to s/2

is
is
1.1
A

1.1. The surface representing the maximum shearing stress rm above the plane of

Fig. 18-35. Fig. 18-36.


Fig. 18-35. Spreading of plastic region from hole in a tension member.
Fio. 18-36. Photoelastic test of transparent plate subjected to pure tension in the direc
tion indicated by arrows and having hole. The dark bands are the pictures of the iso-
a

chromatic lines (of the lines having certain definite colors) as they appeared on the screen.

the plate has two sharp peaks at the points marked = in the figure and situated
k

3
<t>

on the boundary of the hole at = a, = jt/2, and 3n/2.


r

Figure 18-34 reveals several remarkable facts: (1) that the true stress-concentration
effect in an elastic material extends only over a comparatively small area. High values
of the maximum shearing stress are encountered only in the vicinity of the peaks of
the surface representing rnu = f(r,<t>). (2) The contour lines for the value = i.e.,
\,
k

the lines along which the maximum shearing stress the same as that in the undis
is

turbed parts of the plate, extend to infinity. (3) The effect of stress concentration
is

furthermore not uniformly distributed around the hole but starts at the two points

These curves have been calculated by A. M. Wahl and checked by photoelastic


1

tests by R. V. Baud at the Westinghouse Research Laboratories, to whom the author


expresses his indebtedness (see Fig. 18-36) and Sec. 18-4.
292 THEORY OF FLOW AND FRACTURE OF SOLIDS

k = 3 (at the "peaks" of the surface r.u of the hole) and


= f[r,<t>] on the boundary
tends then to progress along a cross-shaped area as indicated by Fig. 18-35. l

18-4. How Plastic Flow Starts around a Hole. According to this we


must expect that in an elastic material with a well-defined yield stress
the first yielding must start at the two points situated on the boundary
of the hole on a diameter perpendicular to the direction of the tension
and at a tensile stress in the plate which is equal to one-third of the yield
stress ffo in pure tension. But this local yielding is very hard to observe
because at first it is practically restricted to two points. As the tensile
stress s gradually increases, yielding spreads and very soon tends to
progress along two comparatively narrow strips symmetrically situated
with respect to the axis of tension and at an angle of about 45 deg with
the direction of tension.
The shape of the contour lines of the surface of maximum elastic
shearing stress was checked by photoelastic tests. A transparent model
of bakelite 2 in. wide having a hole of 0.32 in. diameter was tested in
tension and projected on a white screen using polarized light. The col
ored fringes of the loaded model, which could be observed on the screen,
correspond in the photoelastic test to those points in the plate, where the
difference of the two principal stresses is equal to a certain constant.
Thus the picture of the colored lines on the screen reveals the lines of
constant maximum shearing stress. A photograph of these colored lines
is shown in Fig. 18-36. The agreement between the observed dark bands
corresponding to lines of the same color in the photograph, Fig. 18-36.
and the calculated contour lines in Fig. 18-34 was very satisfactory.
The results of some compression tests with mild-steel test pieces, hav
ing drilled holes, are shown in Figs. 18-37 to 18-39. The test pieces were
approximately prismatical in shape and had, in the middle, a rectangular
cross section of 18 by 24 mm. The surfaces of the specimen containing
the holes were photographed using the Schlieren method described in Sec.
18-1. In the test piece of Fig. 18-37, the hole was perpendicular to the
direction of compression, in that of Figs. 18-38 and 18-39 the hole was
at an angle of 77 deg and 47 deg, respectively. A schematic representa
tion of the production of the layers visible on the surface for the case of a
hole perpendicular and for the case of a hole at an angle of 45 deg to the
direction of compression is shown in Figs. 18-40 and 18-41. With
1 In brittle materials the presence of a hole or notch with sharp corners is always
accompanied by great danger of fracture especially under impact. The effect of a
hole or of a concentration of stress on the fatigue limit, however, seems not to be so
pronounced as one would expect. The hole does not lower the endurance limit so
much as one would expect from the conclusions drawn from the theory of elasticity
mentioned above.
STRAIN AND FLOW FIGURES 293

decreasing angles it must be expected that both slip planes tend to come
together and at an angle of 45 dcg with the axis of the prism they coincide.
This expectation was confirmed by the tests shown in the Figs. 18-37 to

Fig. 18-37. Fig. 18-38. Fig. 18-39.


a = 9, de(?. a =77 deg. a = 47 deg.
Figs. 18-37, 18-38, and 18-39. Flow figures on prisms of mild steel containing holes.
Compression tests, a is the angle of axis of hole with axis of prism.

18-39. In Figs. 18-37 and 18-39, besides the heavy flow lines, a grained
appearance of the surface may be noted. These fine relief markings
appeared during the progress of the test and were probably caused by an
additional bending stress produced by an unavoid
able eccentricity of loading. The "wave crests"
of this fine relief marking, running perpendicular to
the direction of compression, are the traces of the
flow layers caused by the superimposed bending
stresses due to buckling. Moreover, on the highly
polished surface of the rectangle (Fig. 18-2) may Fig. 18^0. Fig. 18-41.
be seen similar markings representing fine flow
layers. These slight irregularities of a well-polished surface which are
barely visible to the naked eye are shown up well by the special method
of illumination described in Sec. 18-1.
Figures 18-42, 18-43, and 18-44 show the development of slip layers as
294 THEORY OF FLOW AND FRACTURE OF SOLIDS

Fio. 18-42. Fio. 18-43.


Figs. 18-42 and 18-43. Flow figures on compression test pieces of steel having hole.

Fig. 18-44. Flow fig- Fio. 18-45. Cross of slip


ures on mild steel. layers produced by hole in
compression test piece of
work-hardened copper.
STRAIN AND FLOW FIGURES 295

they progress from a hole under increasing stress, the test pieces being
subjected to compression. The test pieces of low-carbon steel were, after
machining, annealed at 930°C, polished, and tested.
These and other tests showed that the softer a polycrystalline metal is
(i.e., the more gradual the transition of the elastic portion of the stress-
strain curve to the plastic portion), the less sharply are the slip layers
marked. Test pieces of mild steel having a sharp peak or sharp break
in the tensile stress-strain curve with a well-defined yield point show up
the two slip layers the best. On the contrary, a soft material like an
annealed copper will not show sharp slip layers.

Fio. 18-46. Hole in region subjected to pure shear.

If this point of view is correct, it should be expected that the phe


nomenon of the formation of Luders' lines, so characteristic of low-carbon
steel and, for example, not known in tests with copper, should be also
produced in materials other than low-carbon steel. This was confirmed.
The test piece shown in Fig. 18-45 was machined from a copper bar which
was first subjected to severe cold working by compressing the bar about
33 per cent. This initial deformation had the effect of cold working the
copper and bringing it artificially into a condition mechanically similar
to the annealed steel, i.e., both metals showing a well-defined yield point.
The specimen of hard copper with a hole showed the cross of slip layers
quite in the same way as the steel test pieces described before.
We must therefore conclude that the formation of Luders' lines in
steel, or of similar bands and plastic layers in other metals, must be
expected if the material has a well-defined yield point, a slight concentra
tion of stress helping to start or develop these single plastic layers.

18-6. Hole in Region Stressed in Pure Shear. It is of interest to determine the


distribution of the lines of constant maximum shear in a plate with a hole, the plate
being subjected to pure shear, i.e., a uniform tension combined with an equal com
pression acting at right angles (see Fig. 18-46). In such a case, the distribution of
29G THEORY OF FLOW AND FRACTURE OF SOLIDS

the lines of constant maximum shear have the form shown in Fig. 18-47. These may
easily be obtained by superimposing a distribution of stresses represented by Fig. 18-34
upon another set of stresses similar to those represented by Fig. 18-34 but displaced
through an angle of 90 deg. The superimposed stresses are to be taken as negative;

Direction of tensionStress

Flo. 18-47. Elastic distribution of stress around hole in region subjected to pure shear.
The contour lines indicate the curves of constant maximum shear.

the result is the series of contour lines shown in Fig. 18-47. These represent the locus
of points having constant values of k = 2rm«I/s as before.
It is perhaps worth mentioning here that such a stress distribution around a hole
might be obtained in the web of an I beam, subjected to a severe transverse shearing
load. The central part of the web can then be considered as stressed in pure shear, and
if the amount of shear is large enough, the material should yield in the vicinity of a
hole in the web along two perpendicular planes.
CHAPTER 19

THE YIELD POINT OF MILD STEEL. PLASTIC FRONTS.


OBLIQUE FRACTURE IN FLAT BARS
Much has been written about the peculiar behavior at the beginning of
yielding and regarding the shapes of the transition curve from the elastic
to the plastic range of strains in the tensile stress-strain curve of low-
carbon steels. The German Society of Steel and Mining Engineers
instituted careful investigations on the mechanical and metallurgical
factors influencing the shape of the transition curve.1 Among the investi
gators who have studied these factors Gilbert Cook,2 J. L. M. Morrison,3
M. Moser,4 F. Korber,5 H. Quinney," G. Welter,7 E. Siebel and S. Schwai-
gerer,8 Miss C. F. Elam,9 J. Winlock and R. W. E. Leiter,l0C. W. Mac

Cf. various reports by the WerkstofTausschuss des Vereins deutscher Eisenhutten-


1

leute and in Stahl u. Eisen, Dtisseldorf.


■ The Yield Point and Initial Stages of Plastic Strain in Mild Steel, Trans. Roy.

Soc. London, ser. A, vol. 230, pp. 103-147, 1931; also Engineering, London, Sept. 11,
1931, p. 343.
" The Yield of Mild Steel . . . , Proc. Inst. Mech. Engrs. London, vol. 142, pp. 193-
221, 1940.
4 Uber die Elastizitatsgrenze und Streckgrenze, Werkstoffausschussber., No. 96,
Verein deutsche Eisenhlittenleute, Dtisseldorf, 1926; also M. Moser, Grundsatzliches
zur Streckgrenze (C. Bach Festschrift), Mitt. u. Forschungsarb., No. 295, p. 74, 1927,
Berlin; and Hartemessungen auf Fliessfiguren, Stahl u. Eisen, vol. 48, p. 1601, 1928.
'Uber die mech. Eigenschaften u. das Geftige kristisch gereckten u. gegliihten
Weicheisens, Mitt. Kaiser-Wilhelm-Inst. Eisenforsch., vol. 4, pp. 31-60, 1924. Also
"Das Problem der Streckgrenze," International Congress for Testing Materials,
September, 1927, Amsterdam; and various reports issued by the Kaiser-Wilhelm-
Institut fur Eisenforschung, Dtisseldorf.
•Time Effect in Testing Metals, Engineer, vol. 157, p. 332, 1934; also vol. 161,
p. 669, 1936.
7 The Upper and Lower Yield Points and the Breaking Strain, Melallurgia, vol. 13,
p. 183, 1936. Also Metallwirtschafl, vol. 14, 1935. Welter, G., and S. Gockowski,
Some Fundamental Factors Regarding the Stress-strain Diagram for Mild Steel,
Melallurgia, vol. 18, p. 99, 1938.
* Die Streckgrenze beim Zugversuch, No. 393, Arch. EisenhiU-
Werkstoffausschussber.
tenw., vol. 11, p. 319, 1937-1938.
• The Influence of Rate of Deformation on the Tensile Test with Special Reference
to the Yield Point in Iron and Steel, Proc. Roy. Soc. London, ser. A, vol. 165, pp. 568-
592, 1938. In this paper many further references to work on the yield point of mild
steel may be found which could not be quoted in the text.
10 Some Factors Affecting the Plastic Deformation of Sheet and Strip Steel, Trans.
ASM, vol. 25, p. 163, 1937. Also Trans. ASME, vol. 25, p. 163, 1937. Also Trans.
ASME, December, 1938.
297
298 THEORY OF FLOW AND FRACTURE OF SOLIDS

Gregor,1 E. A. Davis,2 M. J. Manjoine,8 and J. Miklowitz4 may be


mentioned besides those whose papers were quoted in the previous
chapter.
19-1. The Peak of the Stress-Strain Curve of Mild Steel. Upper,
Lower Yield Stress. Yield Point Elongation. As was mentioned in
Chap. 3, mild steel bars when tested in tension under an increasing load
first stretch perfectly elastically until the load
smess' x° reaches a certain value at which the permanent
deformations suddenly start to develop. If
round bars with a well-machined or polished
Is ISfc surface are tested on the usual type of tensile
'■ testing machine and if the two connecting por
tions at the ends of the cylindrical gauge length
<c —£». to the heads of the test specimen are gradually
strain enlarged (have a very large fillet radius), it is
Fio. 19-1. The peak of freqUently observed that the pointer of the
the stress-strain curve. , . , , .
force indicator of the testing machine, after
it stopped for a short instant when the yield point of the steel has
been reached, suddenly reverses and that the load drops rapidly, often
10 per cent or more (Fig. 19-1) while the first flow layer may be detected
on the specimen. C. von Bach,6 who first paid attention to this phe
nomenon, introduced the terms "upper" and "lower" yield point in 1904
for the maximum load at which yielding starts and for the second smaller
load under which the bar continued to stretch for a certain time. An
upper and lower yield point is particularly characteristic of normalized
low-carbon steels (steels that have been annealed at a temperature
slightly above the transformation temperature). The drop of the load
at the yield point may occur but may not be so pronounced in flat bars;
it can be suppressed by making the fillets at the ends of the cylindrical
portion of a round bar very sharp. This was shown in tests by the author
by using tensile test bars having a groove in circumferential direction of

1 The Yield Point of Mild Steel, Trans. ASME, June, 1931. The Formation of
Localized Slip Layers in Metals, Metals & Alloys, February, 1933. The Tension
Test, Proc. ASTM, June, 1940.
1 The Effect of the Speed of Stretching and the Rate of Loading on the Yielding

of Mild Steel,J. Applied Mechanics, vol. 5, December, 1938.


* Influence of Rate of Strain and Temperature on Yield Stresses of Mild Steel, J.
Applied Mechanics, December, 1944.
* The Initiation and Propagation of the Plastic Zone in a Tension Bar of Mild
Steel (a, under eccentric loading and 6, as influenced by the speed of stretching),
J. Applied Mechanics, ASME meeting, December, 1946.
* Zum Begriff der Streckgrenze, Z. Ver. deut. Ing., p. 1040, 1904; p. 615, 1905.
YIELD POINT OF MILD STEEL 299

different proportions (see Figs. 19-2 and 19-3). x While on round bars
with a highly polished surface and with a gradual increase in diameter at
both ends an upper yield stress a\ = 3,387 kg/cm2 and a lower yield
stress c% = 2,969 kg/cm2 (mean values of seven tests) was observed, on
bars having a sharp constriction according to Fig. 19-2, the values
<ri
= 3,676 kg/cm2 and <rj = 3,612 kg/cm2 were recorded. Furthermore,
it was shown by means of a Martens prism extensometer that the first
signs of a permanent strain in the case of the bars with the sharper con
striction could be observed at stresses well below 3,000 kg/cm2. At a
first glance it would appear that a short constriction tends to increase the
yield stress. As the last figure shows, just the
contrary is the case, because of the high concen- I

Hi
tration of the stresses within the elastic range of

[mm
strains through which localized yielding starts at
much lower mean stress. The higher values for a\
and at in this case were observed because a short
constriction prevents the full formation of a flow ,,, , , , ,
i • i- FlG- 19~2- Fl°- 19_3-
layer across the section and because the free lateral Fios. 19-2. and 19-3.
Grooved bare for tensile
contraction in the neck which must accompany an
axial extension cannot develop.2
As Moser, Ludwik, and others have pointed out, the lower yield stress
is less apt to be influenced through the shape of the test specimen; the
upper yield stress, on the other hand, can be raised by carefully avoiding
those disturbances which cause a concentration of the stresses in or near
the ends of the cylindrical portion of the specimen. Thus more practical
significance has been attributed to the lower yield stress as a strength-
determining quantity. As long as the horizontal portion BC (Fig. 19-1)
in the stress-strain curve of a mild steel is recorded, the test bar does not
yield uniformly. The permanent strain corresponding to point C is called
the yield point elongation. After the curve starts to rise steeply again
(branch CD) the bar stretches uniformly. By inserting an elastic spring
in series with the test specimen and by varying its stiffness, Siebel and
Schwaigerer have shown that the slope under which the curve AB drops
can be reduced (AB') by decreasing the constant AP.Ax of the elastic
spring. The steepest slopes in the drop of the load are obtained in bars

'Dissertation, Berlin, 1911.


* It
is well known from early experiments by Kirkaldy, Martens, Rudeloff, and
others that the "ultimate strength" (maximum stress, referred to the original mini
mum cross section) of a cylindrical bar having a short cylindrical constriction like
the one represented in Fig. 19-2 or 19-3 is considerably higher than the ultimate
strength of a round bar because of the prevention of normal necking in the constricted
portion.
300 THEORY OF FLOW AND FRACTURE OF SOLIDS

tested in machines which respond with very little relative movement


between the two heads upon increase of the load in the absence of a test
specimen.

Unit Elongation
Fio. 19-4. Short-time tensile tests with steel at elevated temperatures. (According to
"
C. Bach and R. Baumann, Festigkeitseigenschaften und Gefilgebilder der Konstruktions-
materialien," 2d erf., p. 11, Verlag Julius Springer, Berlin, 1921.)

According to tests by Bach and Baumann and others, the phenomenon


of an upper and lower yield stress of mild steel disappears gradually with
a rising temperature (Fig. 19-4).
19-2. Plastic Front in Tension Bar.
Tests with Nylon. Before further obser
vations on the discontinuous yielding of
mild steel are reported, an interesting
example in an entirely different type of a
material, in whose inner structure a sudden
change under a definite load also occurs,
will be mentioned. It came to the atten
tion of the author when in 1938 some
A tests with thin filaments of nylon were
watched and it was noticed that they
started to yield in tension in a manner
strongly reminiscent of that of bars of
mild steel. A grade of nylon in which
the long molecule chains are oriented at
A U random exhibits this behavior. When
the load is increased, under a certain load
a deep constriction is seen suddenly to
^^^^^^^^ form which increases in length while its
Fio. 19-5. Sharp constrictions two ends move in opposite directions like
forming in nylon filaments
stretched beyond the yield point.
two congruent wave fronts at low speed
along the filament and the load remains
unchanged. Two examples of such plastic fronts are shown in Fig. 19-51
after the yielding was interrupted.
1 From X-ray studies by the du Pont Company, Wilmington, Del., it is known that
the molecules of nylon (a high polymer) in the contracted portion of a filament
YIELD POINT OF MILD STEEL 301

A most surprising fact became apparent in these tests, namely, the


enormous "yield point elongation" which could be observed in these
filaments, amounting to 200 to 300 per cent while the yield point elonga
tion of a mild steel is of an order of 3 to 5 per cent.
Although the molecular mechanisms causing the discontinuity of the
stress-strain curve at the yield point in these two materials are very differ
ent, it seemed of some interest to compare their mechanical behavior at

Fio. 19-6. Working lengths in nylon.

the yield point. High polymers such as nylon deform at higher stresses
when they are stretched at faster rates. Let A denote the area of cross
section, « the strain, and a the stress in the distance x in the filament, and
let the stress be <r0 in the original area A 0 and P0 = A 0ao the load at which
the double funnel-shaped constriction (Fig. 19-6) has fully developed.
(It is assumed that the constriction starts to form at a lower load. Before
o"o is reached the stress-strain curve must gradually depart from its
straight portion corresponding to the elastic strains.) In the contracted

rearrange themselves in the direction parallel to the fiber axis. The formation of the
constriction and the stress-strain curves of nylon were studied by J. Miklowitz at the
Westinghouse Research Laboratories for the National Defense Research Committee
and reported by him in two papers in J.
Colloid Sci., vol. 2, No. 1, pp. 193-215, 217-
222, February, 1947. The results given in the text above are taken from the first
paper.
302 THEORY OF FLOW AND FRACTURE OF SOLIDS

portion under the load P0 = A&o the true stresses a increase toward its
minimum section A \ to the value <n and

P0 = Ao<t0
= Aa = A i«n . (19-1)

Half the length of the fully formed constriction may be called the
"working length" I in the filament. When the total length of the fila
ment further continuously increases, the load remains approximately at
the constant value P0, while new portions of the fiber are yielded and the
structure in them changes from the random to the oriented structure. A
later stage of the stretching process is shown in the lower sketch of Fig.
19-6.
Since the volume does not change during the stretching we have

A(l +e) = A,(l + «,) = A0. (19-2)

Thus, to each section A in the neck and local strain t corresponds a true

j
stress

<r = = <r0(l + e) (19-3)

and we see that a must be a (linear) function of the local strains e in the
"working length" I. This is reminiscent of the strain hardening of a
ductile metal. The true conditions are somewhat more complex, if the
dependence of the true stresses on the rates of strain is also considered.
While the two working lengths I move along the fiber, as two waves in
opposite directions, those portions which have not yet yielded, as well as
the yielded ones, must further increase in length and contract in their
sections by small additional amounts because of the "viscous" behavior
of the nylon. This will become more pronounced in the tension tests at
very slow rates and will have the effect that the profile of the plastic wave
moving along the filament is stretched out more and more as the test
lasts longer. In consequence of the viscosity of the material the load P0
cannot remain constant but must drop slightly because the "true" local
rates of strain gradually decrease if the two ends of the filament are
moved with a constant velocity relative to each other, as was assumed in
these tests, and the length of the fiber increases two- to threefold.
In the lower set of curves in Fig. 19-7 the stress <r0 (the load divided by
the original area of the cross section of the filament) is plotted as a func
tion of the mean conventional strain (computed by dividing the relative
displacement between the two clamps to which the filament was attached
by their original distance). Four tests are represented for four given
constant velocities with which the head of the little tensile machine in
which the filaments were stretched was moved. It should be noted that
YIELD POINT OF MILD STEEL 303

it is not very sensible to speak of a mean strain in these tests since the
axial strain varies within the working length I theoretically between
the values zero and <i. The abscissas in Fig. 19-7 have a good meaning,
however, if they are interpreted not as the mean but as the local axial

/ 2 3
MEAN CONVENTIONAL STRAINS
Fig. 19-7. Tensile Btress-strain curves of nylon filaments. (Testa by J. Miklowitz.)

strains « of the actual cross sections A within the working length I of the
plastic zone in the filament. According to this second interpretation, we
may compute in each cross section area A the true stress o- = o-0(l + «).
a is represented in Fig. 19-7 by straight lines for each of the four tests
304 THEORY OF FLOW AND FRACTURE OF SOLIDS

which must radiate from the point 0 (having the abscissa t = — 1) to


the left of the origin O2 of the true stress <r local strain t curves.1 This
has been shown in the upper set of "true stress-strain curves" a = f(t)
in Fig. 19-7.
From the true stress-local-conventional-strain curves of Fig. 19-7 two
important conclusions may now be drawn concerning this discontinuous
yielding in a material. First, we may state that discontinuous (local)
yielding must occur in a material if the true stress-conventional-strain curve
a = starts with that critical slope da/dt = which is just equal to the
/(«) a0
yield stress o-0 (or with a slope da/dt which is smaller than <r0)- For materials
having this property the test specimen can stretch under constant load
(stress) tr0 (or a decreasing stress <r0). Second, the yield point elongation
is a highly speed-dependent phenomenon. It increases in the case of nylon
filaments 2 in. long from e = 1.04 to « = 2.96 (nearly threefold) over a
range of the head speeds v = 0.004 to 8.7 in. /sec. A third observation
may be added: the faster the nylon fibers were stretched the shorter and
"
sharper the conical transition portion (the "working length) became between
the yielded and the unyielded portions of the filament.2
There are two important differences, however, between the modes of
1 It should be noted that because of the comparatively long working length I of
nylon the stresses a may be approximately computed by dividing the load at the
yield point Pa by the areas A. The thin dotted straight lines radiating from point 0
in the upper figure (Fig. 19-7) are true stress-strain lines which were computed from
the total value of the observed yield point elongations for four additional nylon fibers.
The termination points of these four inclined straight lines indicate the yield point
elongation strains. The largest of them, corresponding to a head rate of v = 8.70 in./
sec (filament F9), was e = 3 (300 per cent).
It is to be noted from Fig. 19-7 that instead of one horizontal branch in two of
the curves <r0 = fit) corresponding to the two fastest speeds of stretching two distinct
horizontal branches were recorded; also that the second branches, which developed
later during the test, were nearer to the strain axis than the first ones. The cause of
this unexpected anomaly was easily located, after it was found that in the latter two
cases not one but two distinct constrictions had formed in the stretched fibers. Since
nylon shows a pronounced effect of the speed of straining, the yield point load dropped
after the second constriction had formed, because in it two new working lengths had
been added and the strain rates in each of the four working lengths I were conse
quently reduced to half of their original values. Double necking may be caused
through nonuniformity of the cross sections or in the properties of the fibers.
It may furthermore be worthy of note that while in mild steel the stress-strain
curve after the yield point elongation (point C in Fig. 19-1) rises quite steeply (branch
CD in Fig. 19-1) the corresponding portions in the lower set of curves in Fig. 19-7
for nylon arc flat and rise quite gradually, also that they appear shifted to the right
in Fig. 19-7 (in contrast to curves for mild steel recorded at increasing seta of strain
rates, which would all practically coincide in the same branch CD of the load versus
strain curve).
8 As will be seen later, the contrary is true in flat bars of steel, in which the length

of the working zone increases with the speed of stretching.


YIELD POINT OF MILD STEEL 305

discontinuous yielding in nylon and mild steel to which attention must


be called. One is due to the large difference in the moduli of elasticity
of these two materials: nylon has a modulus of elasticity in the unoriented
fiber E = 7 X 104 and 2 to 4 X 106 lb/in.! in the oriented fiber, mild
steel E = 3 X 107 lb/in.2 This must have a profound effect on the
initiation of the first constriction (in nylon) and of "the first flow layer"
in steels, respectively. The former may develop gradually, under increas
ing stress, the second frequently under an abrupt drop of load. The
formation of the constriction in nylon at the yield point may be compared
with the process of the gradual necking of a round bar of a ductile metal
just before it breaks (incidentally, the rule da/dt ^ a0 determines the
instant at which the nonuniform mode of stretching and the necking
develops in metals also, as was mentioned in Chap. 8). The state of
stress within the conical necked portions in a nylon filament or in the
neck of a round bar of a ductile metal after the "ultimate strength" is
reached is a rotationally symmetric state of tensile stresses. The state
of stress and strain, however, is quite different within the first layer of
yielded metal and subsequently, after a working zone developed in a mild
steel bar during the process of the yield point elongation, as will be
discussed in the following section.
19-3. Instability of Tensile Equilibrium in Compact Bars at Yield
Point. Suppose that a small region in a round bar of steel has a slightly
lower yield stress than the surrounding portions in the bar. Were this
small region free in space, it would elongate permanently when the yield
stress is reached by a few per cent strain in the direction of the tensile
stress and contract in all lateral directions by half of this amount. Since
it
is,

however, surrounded by material in which the resistance against


yielding greater, permanent strains of the magnitude just mentioned
is

cannot develop in the small region. Suppose that the yield point elonga
tion per cent. The small element of material should contract by
is
4

per cent in the lateral directions. But the surrounding material can
2

deform only elastically since its yield point has not yet been reached.
Assuming that the yield stress equal to 30,000 lb/in.2 and that the
is

modulus of elasticity = 30,000,000 lb/in.2 this strain 0.001 in


E
is

is

axial and 0.0003 in lateral direction (assuming that Poisson's ratio


is

v = 0.3). This elastic lateral strain in the material surrounding the small
yielded region only three two-hundredths or 1.5 per cent of the -perma
is

nent lateral strain in the small region the latter was permitted to deform
if

freely. Apart from very small permanent strain, the latter region will
a

therefore only be deformed elastically.1 This may be the case in many


The case of disturbance caused in an elastic material in which soft or plasti-
1

cized region in the shape of a small sphere has just formed has been analyzed by G.
I.

Taylor (Faults in a Material Which Yields to Shear Stress While Retaining Its
300 THEORY OF FLOW AND FRACTURE OF SOLIDS

other weaker regions. Consequently the mean stress can increase above
the local yield values, and thus the internal equilibrium of the stresses
in the bar will gradually become unstable if the load is further increased
and provided that no sharp concentrators of stress are present, such as
sharp fillets at the ends of the cylindrical portion, or small holes or
notches. At a certain higher load, however, a new type of permanent
deformation becomes possible which can develop without a lateral contraction,
namely, a simple permanent plane shear within a thin layer of the bar inclined

Fig. 19-8. Equilibrium of stresses in a thin flow layer in compact specimen of steel.

at an angle of 45 deg with respect to the direction of tension. In Sec.


\5-\3d it was shown that steel under a simple shear yields under a shear
stresst = <r0/\/3 = 0.577o-0 where <x0is the "lower" yield stress of steel in
simple axial tension. This plane state of a simple shear t in a thin layer
A BCD (Fig. 19-8) of material which is inclined at an angle of 45 deg with
respect to the direction of tension <r has the principal stresses v\ = t,
oY = — t and as = 0 (in the directions indicated in Fig. 19-8). Such a

Volume Elasticity," Proc. Roy. Soc. London, vol. 134, p. 1, 1934), who assumed that
in a homogeneous field of stress a soft or weak region which is ready to yield is present
in the shape of a small "fluid" sphere and interesting conclusions were drawn by him
concerning the mode in which such nucleuses would cause either spreading of the
plastic deformation into the surrounding elastic material or initiation of a crack.
Related cases (holes present in stressed material) will be treated in Chap. 20 and in
Chaps. 36 and 37 where the initiation of a flow layer by a semicircular groove in a
body stressed by simple shear will be treated.
YIELD POINT OF MILD STEEL 307

plane state of simple shear in a flow layer A BCD in a tension bar, how
ever, cannot exist for itself since it would subject the elements of material
in the layer to a compression stress <rV in the direction perpendicular to
the tension stress <r.

By superposing on the state of simple plane shear <r\ = — <r2' = t,


<r3'
= 0 a state of equiaxial tensions <ri" = <r2" = <r3" = t the resultant

principal stress <r2 = a% + <r2" = — r + t = 0 may again be made to


vanish without an effect on the condition of plasticity. The resultant
principal stresses in a material element situated within the flow layer
will be

(19-4)

The tensile stress a in the bar, required to produce in the plane AB of


maximum shear a normal and shear stress equal to t, is equal to twice this
value, however, and hence we see that the round bar of steel will start
to yield in a thin layer inclined at an angle of 45 deg with respect to the
direction of the tension under an "upper" yield stress

a = 2t = = 1.155<ro
(19-5)

which is 16 per cent higher than the "lower" yield stress <r0 for unrestricted
yielding in tension, and we note that the metal within the flow layer is
subjected to tensile stress (<t3 = t) in the normal direction to the plane of
Fig. 19-8. These stresses <n are caused by the constraint exerted by the
two adjoining elastic portions of the bar not yet participating in the
plastic deformation.1 We may expect that these idealized conditions
will seldom be encountered in the purity in which they were here assumed
and that the upper yield stress may not reach the high value which was
just computed. If the steel is susceptible to strain aging while it yields,
this may obscure the mechanical causes for the delay of the beginning of
pure uniaxial yielding in compact specimens of steel, and perhaps even
much higher values of the upper yield stress may be observed in bars
with circular or square cross section.* In flat bars of small thickness
other conditions prevail, to which reference will be made below.

1 These normal stresses <r3 in the flow layer must create certain shearing stresses
which must be concentrated along the circumference of the ellipses bounding the flow
layer which constitutes a secondary disturbance in the system of stresses responsible
for the plane strain in the layer.
2 By using the
plastic stress-strain relations for the plane homogeneous state of
stress prevailing in a thin flow layer the boundary conditions for the mean strains in
308 THEORY OF FLOW AND FRACTURE OF SOLIDS

We may state, in conclusion, that the initiation of plastic strains — of


a tensile nature — of several per cent in the nucleuses of slightly weaker
material is retarded in compact specimens, since these nucleuses are
surrounded by elastic material which is unable to accommodate such
large strains and thus an unstable elastic equilibrium of the internal
stresses must develop in smooth cylindrical bars till the uniform stress
has increased to that value under which the second mode of distortion
becomes possible through localized yielding in a thin layer inclined along
a plane of maximum shear, at least temporarily, for a short impulse of
flow.1 After the slip band has formed and this layer increases in thick
ness the metal within the plasticized layer will be drawn in radially (under
axial tension) or be squeezed out in the radial directions (under axial
compression) performing somewhat as a layer of viscous material would
do between two rigid plates. This mode of yielding appears on the sur
face of the specimens in some of the prisms of square cross section of which
photographs are reproduced in Chap. 18. (See the formation of shallow
grooves in tension and of ridges in compression tests explained there.)
As this layer increases further in thickness the constraint originating
from the undeformed portions of the bar gradually becomes weaker and
the stresses throughout the working layer are redistributed. This
explains the drop of load from the upper to the lower yield stress. It
may be added that in the boundary surface of the advancing plastic zone
after it has increased in thickness still some stresses along the edges of the
surface will remain higher than their mean value so that yielding will
progress even though the total load has now been reduced to its lower
yield point value.
Two additional facts influence this second mode of yielding while the

the two parallel oblique plane sections AB and CD (Fig. 19-8) may be considered.
These relations lead to the same distribution of the stresses as the one which was just
discussed (see author's paper in Proc. Inst. Mech. Engrs. London, p. 155, 1947). By
superposing on the state of simple shear triaxial equal tensions for which

<Xl = <72 = <rj < T

a certain residual compression stress oi may be retained in the flow layer and a smaller
will not be possible to satisfy the
it,

tensile stress a% obtained perpendicular to but


it

conditions for the mean normal strains «, .„ «„ = «j = in the flow layer parallel
(x
0

to the direction of slip, perpendicular to the plane of the flow layer; in the direction
y

of <rj) required for this case of flow.


Discontinuous yielding in a thin layer will not be induced the stress <rt>(the load
if
1

on the bar) increases with the strain t. This the case when d<r/dt > <ro. This occurs
is

during the stretching of soft ductile metal like copper or after an interrupted test
a

and renewed further stretching, in contrast to the behavior of mild steel at sharply
a

defined yield stress <r0 = const.


YIELD POINT OF MILD STEEL 309

yielded zone grows under a constant load, which have not yet been con
sidered, namely, (1) the strain hardening of the metal combined with (2)
a simultaneous decrease of the cross-sectional area. The rear end of the
advancing "working" zone has a smaller cross section and higher mean
stresses than its front portion. This zone in its fully developed stage
may well be observed on flat bars which have partly yielded. Conditions
in it are reminiscent of the ones which have been described for the necked
shoulders in the nylon filaments, although the distribution of the stresses
in the working zone in a steel bar cannot be so simple as in the necked
portion in a nylon filament because the length of the "working" zone in a
steel bar in proportion to its diameter is much shorter than in nylon.

C-STRAIN STRAIN

(a) (b)
Fig. 19-9. Two types of stress-strain diagrams for mild steel. (a) Constant head rate
test. (6) Constant load rata test. «i yield point elongation.

One additional circumstance contributes to the formation of a sharp


peak in the stress-strain curve of mild steel. During the increase of the
load the heads of the testing machine move relative to each other with a
given velocity. If the bar started to yield continuously over its cylin
drical portion, this speed would determine the mean permanent rate of
strain in the bar. If yielding, however, starts within a very thin layer
by shear and the entire "working" length is reduced to the dimensions
of the thickness of this flow layer, i.e., to a very small length, the perma
nent strains in it must be produced under very high rates of strain or shear.
This is an additional reason for an increase of the load producing the
"upper" yield point. Some tests on this increase of the yield stresses
with the rates of strain are reported in the following section.
19-4. Increase of Yield Stress and Yield Point Elongation with Rate
of Strain. The slope of the curve describing the drop from the upper
to the lower yield point in the stress-strain curve of a soft steel depends,
as the tests by Siebel, Welter, and Miklowitz have shown, on the elastic
ity of the testing machine. The more rigidly the stationary head of the
testing machine reacts upon an increase of load the steeper is this curve.
A "lower" yield point will not be observed, however (Fig. 19-96), if the
310 THEORY OF FLOW AND FRACTURE OF SOLIDS

steel is tested in a machine constructed like the one of which Fig. 19-10
shows a view. This machine permits the stressing of a tensile specimen
with a constant if the heavy weight which runs
rate of increase of the load
on wheels on the horizontal lever arm is moved with a constant velocity.
A set of stress-strain diagrams recorded by E. A. Davis for a mild steel

Fig. 19-10. Constant load rate testing machine.

obtained by running tensile tests on this machine by varying the rate of


loading is reproduced in Fig. 19-11 showing that no lower yield stress
appeared in these curves and that the yield stress and yield point elonga
tion increased with the rate of loading. A ten thousandfold increase
of the rate of loading raised the (upper) yield stress for one mild steel by
30 per cent and the yield point elongation increased four times. Since
this machine does not permit a drop of load (as long as the tests are not
run with such high rates of loading that a static equilibrium ceases to
exist) and yielding starts under the high values of the upper yield stresses,
the corresponding rates of plastic strain in the "working" lengths of the
YIELD POINT OF MILD STEEL 311

specimens must in general be very high compared with the rates of strain
attainable in the usual type of testing machine. This explains the con
siderable relative increase of the yield stress which was observed when the
rate of loading was increased.
The influence of an increase of the rate of straining on the phenomenon

_A_

V
c/

7E/1
/Ff

Cl
5

t
497 LBS

\
STRAIN
Flo. 19-11. Recorded stress-strain diagrams from constant load rate testing machine for
mild steel at room temperature. Hound bars: 0.355 in. diameter, 2 in. gauge length.
Rate of loading, pounds per square inch per second: A = 6,500, B = 1,630, C = 264,
D - 61.1, E = 7.03, F -0.29. (Teats by E. A. Davis.)

of yielding of mild steels was studied by various investigators. We men


tioned some earlier work in Chap. 3, page 21, leading to the conclusion
that the yield stresses a of ductile metals increase as the logarithms of
the rates of stretching v = dt/dt

a = Ci + Ci In (19-6)
I/O

Winlock and Leiter subjected small flat bars of the shape shown in
Fig. 19-12 of five grades of low-carbon steels (0.05% C) of the type used
for deep drawing on a specially built small testing machine to tensile
teats under given constant rates of strain. The results of their tests for
one of the steels are reproduced in plots of load-deformation curves in
312 THEORY OF FLOW AND FRACTURE OF SOLIDS

Fig. 19-12 showing that the yield stress and the yield point elongation
increase with the "mean" rates of strain (their values are given in the
legend to Fig. 19-12). Their results corroborate the observations by
Davis in the lower range of rates of strain of the order 0.002 to 4.44 1 /min ;
the function <r = /(f) which they observed deviated perceptibly from a
straight line in a a, In v plot. Very little indication of the presence of an
"upper" yield point was disclosed in these tests, which is in agreement
Load
lb. Stress
soo\
4X,,*'±,
in*
too

4,0

too
, Strain e
+1 U-i

,.3,,
\+-t.ts" Wo.,3*"

ASTU Standard

\
small flat bar
Test No. i t s 4 s e 7 e
S
Rate of °-,,' \
°-,0e ,,st ojes o soVX15,
o,"v^-^v--^.v~~-^ /
Strain min.
Fio. 19-1: Tension tests by J. Winlock and II. W. Leiter (1936) on effect of rate of strain
on yield point elongation of a 0.04 % C steel.

with other observations according to which tests on flat bars usually do


not show such a pronounced upper yield point as round bars.
The dependence of the yield and flow stresses in low-carbon steel on
the rate of straining was carefully investigated by M. J. Manjoine1 over
a wide range of rates of strain 10-6 to 103 per sec corresponding to tensile
tests of a duration from the order of 24 hr for the slowest to a few thou
sandths of 1 sec for the fastest tests run for stretching a specimen at a
uniform rate until the point of fracture.

1 The small round steel specimens of 0.2 in. diameter and 1 in. gauge length were

annealed for 1 hr at 920°O and tested at small rates in a constant strain rate screw-
driven tensile machine permitting to stretch the specimens under given head rates.
The fast tests were made in the high-speed machine which was reproduced in a photo
graph in Fig. 3-10. For further details see the papers quoted in the footnote on p. 23
in Chap. 3. Only the results of tensile tests made at room temperature and at 200°C
corresponding to the so-called blue heat range of mild steel are reported here.
YIELD POINT OF MILD STEEL 313

Two sets of stress-strain curves obtained at room temperature and at 200°O are
reproduced in Figs. 19-13 and 19-14 from which the influence of the rate of strain on
the tensile properties, on the lower yield stress <r„, on the ultimate strength <r„ (stress
at maximum load referred to original area of cross section of bar), and on the total
elongation in per cent was determined. These quantities and the ratio <ry/au are
represented as ordinates in Figs. 19-15 above the rates of strain dt/dt (« conventional
strain) plotted on a logarithmic scale as abscissas. It is seen from Fig. 19-15 that the
lower yield stress er„ increases much faster than the stress au at the maximum load with
the rates of strain and that the curve representing <r„ has a flat dip. This dip is due
to "strain aging" known to occur in low-carbon steels even at room temperature.
Under strain aging the change of the physical properties (an increase of yield stress
after a previous cold deformation and a rest period followed by an exposure to a
slightly elevated temperature) is understood. Furthermore, the stress-strain curves
in Fig. 19-13 show that the yield point elongation increases in a pronounced way with
the rate of strain. The true flow stresses
a = <r0(1 + t) at given values of the strain «
(assuming that <r0 is the load divided by
the original area of cross section) are repre
sented in Fig. 19-16 in several curves for
the room temperature tests,1 disclosing a
somewhat anomalous behavior of mild steel
concerning the dependence of the flow
stresses on the rates of strain (in the upper
curves), when compared with the normal
flow stress-rate-of-strain plots of nonaging
ductile metal, e.g., copper. (Compare the
corresponding curves which were repro
duced in Fig. 3-10 for copper.) Whereas
the true flow stresses a as a rule increase
with the rates of strain dt/dt, the anoma
lous behavior of mild steel is characterized
by a decrease of a with dt/dt at the lower
rates of strain after the strains t have
reached values beyond 8 per cent or more
in Fig. 19-16. This becomes much more
pronounced in the tension tests run at
200°C reproduced in Fig. 19-17, in which
the humps in the <rcurves in the left portion of the figure at the lower rates of stretch
ing are quite marked after a tensile bar has been stretched by 2 per cent. They are
an indication of the strain aging of mild steel in the blue-heat range. At sufficiently
fast rates of stretching approaching those of impact tests the strain aging becomes
insignificant, the duration of the tensile tests being too short for allowing enough
time for aging, as can be seen from the trend of the a curves in Fig. 19-17 for large
strain rates.
An interesting observation worth mentioning in these tests was that the load in two
of the stress-strain curves in Fig. 19-14 (corresponding to dt/dt = 0.02 and 0.00085
1 During the yield point elongation
<ro = <r„ is constant, but the true stress a increases
with the yield point elongation to. There will be two values of a, namely, a = <r„ at
the beginning and a = <r„(l + *o) at the end of the yield point elongation «0, This
explains the shaded portion between the two curves in Fig. 19-16.
314 THEORY OF FWW AND FRACTURE OF SOLIDS

80
w3.4xlG ■«JK

70

60
__/
'"'
... -r
A" \ IO-^sec.
\\
\
0.5/SEC. \\
j 200 /SEC.-"* \
\40 \
4- 8.5x10 _4■Sec.

$30
I 20

10

-5#
1 10 1 20 I
<? 0 0 <? 0
% STRAIN
Fia. 19-14. Stress-strain curves of mild steel observed at 200°C.

AVERAGE RATE OF STRAIN PER SECOND


Fig. 19-15. Dependence of the tensile properties of mild steel on the rate of straining.
Room temperature tests. (M . J. Manjoine.)
YIELD POINT OF MILD STEEL 315

lOOx

I 20
10

o\ 1 1 1 1 1 1 1 1
l0-e l0-5 ,0-4 /0-j /0-2 io-i , i0i io? io*
RATE OF STRAIN PER SECOND
Via. 19-17. The true flow stresses at various strains in function of the logarithms of the
rates of strain. Mild steel at 200°C.
316 THEORY OF FLOW AND FRACTURE OF SOLIDS

sec-1) after a certain strain (4 and 8 per cent, respectively) was reached started to
oscillate. It is instructive to note that C. F. Elam1 found similar fluctuations of the
load in a quenched aluminum alloy that was tested immediately after quenching,
superimposed on the load-strain curve (Fig. 19-18). The oscillations of the load
started after a certain strain was reached, and their amplitudes increased continuously
during the further extension of the specimen. At the same time the bar became
uneven, proving that the yielding under these circumstances progressed in steps

Fiq. 19-18. Stress-strain curve of Fiq. 19-19. Armco iron tested by Elam.
aluminum alloy observed by Elam. The tensile test was interrupted at inter
vals. The figures refer to the time of rest
after each test in hours.

through the formation of discrete slip bands during each oscillation of the load.
Although the exact conditions under which these oscillations in a stress-strain curve
are produced are not quite clear, there is little doubt that they are the manifestation
of strain-aging in a metal and might have been caused through the true stress-strain
curve taking a critical slope.
A related phenomenon was reported longer ago by P. Ludwik, who observed repeated
peaks in the tensile stress-strain curve of mild steel after interrupting the pulling and
reloading the bar after a certain time had elapsed. Elam found the same in Armco
iron pulled in stages with rest periods (see Fig. 19-19).

19-5. Instability of Tensile Equilibrium in Flat Bars.


Oblique Frac
ture. Although the mode in which flat bars of a metal having a sharply
defined limit of plasticity start to yield varies considerably from steel to
steel and for nonferrous metals, a few interesting observations may briefly
be stated. On certain silicon steel bars (containing less than 0.02 per
cent carbon) or on mild steel bars in which the ratio of the thickness to
the width of the rectangular cross section was small, very sharply defined
thin slip layers were observed at the yield point. If the load acts per
fectly axially and the grips between which the heads are clamped do not
permit additional degrees of freedom of motion, the traces of the first
1 Loc. eit.
YIELD POINT OF MILD STEEL 317

slip layers on the flat sides of the bar are frequently inclined at an angle
of approximately 45 deg (usually slightly larger than 45 deg) with respect
to the tension axis. (See the photographs reproduced in Figs. 18-15 to
18-17.) On steel bars having a larger ratio of thickness to width {e.g.,
M to yi) verV <Awi flow layers are seldom seen to form, but, on the con
trary, a fairly determined plastic "working" portion may frequently be
detectable (Figs. 19-20 and 19-21). In this portion the steel has yielded
between the elastic strain t = 0.001 and the yield point elongation strain

\sm£SS<^

Fig. 19-20. Flo. 19-21. Fig. 19-22. Ten


Figs. 19-20 and 19-21. Observed sile stresses along
working zone and plastic fronts in edge ABCD for flat
flat bars. bar hinged in Oi
and Oi after slip
occurred along EB.

of a magnitude « = 0.03 to 0.05. Corresponding to the latter strains the


area of the cross section at the rear end of the working portion has
decreased by 1.5 to 2.5 per cent and the mean stress is by 1.5 to 2.5 per
cent higher than at the front of the working zone. Assuming that the
grips between which the heads of the flat specimen are very rigidly
clamped cannot rotate around an axis which is perpendicular to the plane
of Fig. 19-21, conditions may be comparable qualitatively to those repre
sented in the photographs of the two rubber models in Figs. 19-23 and
19-24. These photographs should illustrate merely the shapes of the two
plastic working zones at the ends of a flat specimen in their relation to the
elastic zones which are represented in the figures by the wooden blocks
(the angles of 45 deg under which the end blocks were cut are not repre
sentative of the true conditions). A few observations on the working
zones were made by J. Miklowitz on very thin silicon steel specimens
(width 16 mm, thickness 0.35 mm; gauge length 76 mm). He measured
318 THEORY OF FLOW AND FRACTURE OF SOLIDS

At the head speeds, in./min 8.6 X 10-' 2.56 X 10"' 6.16 X 10"'
8 10 17
1.4 1.8 2.8

their lengths in the axial direction and the yield point elongations, as
shown in the table, indicating that in general the length of the working
zones increases with the yield point elongation and with the speed of
stretching.1 The boundary curve
of the plastic front (Fig. 19-21) as
seen on the flat sides of the bar
usually advances as a slightly curved
line which has an average slope of 30
to 35 deg with respect to a line per
pendicular to the tension axis.2
1 Further interesting observations
made
by him were: (1) that the yield stress of
1 per cent silicon steel increased 50 per
cent for a 105-fold increase of the speed of
stretching and (2) that the density with
which the cracks formed in the stress coat
was larger at high than at slow speeds of
stretching, showing that more slip lines or
layers form at high than at low rates of
straining.
2 By introducing a comparatively large
eccentricity for the tensile load (by using
specially designed grips) Miklowitz found
the spreading of the plastic zone in thin flat
bars progressing along wedge-shaped re
gions (Fig. 18-19). This is a case of com
bined tension and bending. It may be
noted that the grips in this case were not
so rigid against a rotation around an axis
which is situated in the plane of principal
bending (in the middle plane of the flat
Fig. 19-23. Fig. 19-24. bars) as in the cases just reported, and
Flos. 19-23 and 19-24. Rubber bands the result was that slip planes showed up
illustrating plastic working lengths in flat as fine lines inclined at an angle of 45 deg
tension bars. The "working lengths" when viewed on the narrow edges of the
are represented in the rubber bands near
the end plates having obliquely inclined specimens.
edges. Flat bars of mild steel rigidly clamped
between hard steel wedges resting on the
flat sides of the heads frequently start to yield in a layer at the end of the prismatic
gauge length. If the enlarged portions of the flat bars start with sharp fillets usually
the first flow layer develops in a direction perpendicular to the axis of the bar. This is due
to the concentration of the stresses near the fillets and the enlarged ends of the
specimens.
YIELD POINT OF MILD STEEL 319

It is well known to testing that wide flat bars or strips


engineers
machined from thin rolled metal sheet when tested in cold worked con
dition in tension do not break in a surface which is perpendicular to the
direction of tension but along an oblique plane perpendicular to the flat
sides of the bar inclined at an angle of approx
imately 55 deg with respect to the axis of the
bar. An example in cold worked steel is illus
trated in Fig. 19-25. The thickness of the bar
decreased along this oblique line while the
width of the bar remained unchanged and the
necking preceding the fracture developed only
in the direction parallel to the smallest dimen
sion of the section.1 The flat bars must have
a ratio of width to thickness of the rectan
gular cross section larger than 6 or 7; other
wise they neck down symmetrically around a
section normal to the bar axis.
F. Koerber and E. Siebel' were the first
ones who called attention to the peculiar way

1 Very strongly cold rolled thin metal sheets have a

pronounced "fiber" structure in the strain hardened


condition and cannot be considered as isotropic mate
rial. It is known that the lattice of the metals having,
for example, a body-centered cubic structure takes a Fia. 19-25. Necking along an
oblique plane in flat bar of
preferred orientation in the crystallites after much cold
cold-rolled low-carbon steel
rolling (one cube face becoming parallel to the plane
(ratio of width to thickness of
of the sheet and a face diagonal of the cube predomi cross section 16) tested in
nantly parallel to the direction of rolling). If flat bars tension. (Teal by J. Aronof-
aky.)
are cut from strongly cold rolled sheets of iron under
different angles with respect to the direction of rolling
and, for example, are tested at liquid air temperature, the fractures preferably select a
direction which coincides with a cubic plane of the reoriented crystallites. This has
been shown in tests with a 4 per cent silicon iron by F. Bitter (Proc. Roy. Soc. London,
vol. 145, p. 668, 1934). The oblique fractures which he observed in flat strips cut at
different angles with respect to the direction of rolling from severely cold rolled thin
silicon iron sheets were influenced in their oblique directions through the anisotropic
structure of these sheets.
R. Hill in a remarkable paper (A Theory of the Yielding and Plastic Flow of
Anisotropic Metals, Proc. Roy. Soc. London, ser. A, vol. 193, pp. 281-297, 1948), after
developing plastic stress-rate-of-strain relations satisfying the conditions of anisotropy
encountered in a flat bar that has been cold rolled, computed the angle for the oblique
straight direction along which a bar of anisotropic metal contracts before fracturing.
1 Zur Theorie der bildsamen Formanderung, Nalurwissenschaften, vol. 16, p. 408,
1928. Also Zur Theorie der Reisswinkelbildung, Mitt. Kaiser-Wilhelm-Inat. Eisen-
forach. Dusseldorf, vol. 10, 1928.
320 THEORY OF FLOW AND FRACTURE OF SOLIDS

in which wide bars of strain hardened metals break in tension tests. (An
incipient oblique fracture in a flat
bar of cold rolled steel tested by them
is reproduced in Fig. 19-26.) They
attributed these types of fractures
to a simultaneous gliding in the metal
on two systems of slip planes, but
their explanation of the phenomenon
(predicting an angle of 55 deg) might
be improved.
Noting that on wide flat bars of
metals having a well-defined yield
point flow layers (Liiders' lines) have
been seen in tension tests which
deviated considerably in their angles
of inclination from the angle of ap
proximately 45 deg observed on com
pact specimens and that such bars
of cold worked metal shortly before
Flo. 19-26. Necking along an ohlique they break contract along a line
plane in flat steel bar tested in tension.
inclined at an angle of 55 to 60 deg
(The steel sheet was reduced in thickness
by 20 per cent by cold rolling before the with respect to the axis of the bar
tensile test was made.) (After F. Koerber in the direction parallel to smallest
and E. Siebel.)
dimension and neck down in a narrow
band along a plane inclined under such an angle, one may search for that
direction with respect to the tension axis in
a bar in which no normal strains need to be
produced under a simple extension.
Suppose that a flat bar of metal having a
well-defined yield point has just started to
yield in a thin oblique layer A BCD inclined at
an angle /3 with respect to a plane perpendic
ular to the axis of the bar (Fig. 19-27). Let
a and t be the normal and shear stress in
plane AR in the elastic zone adjoining the
PLASTIC

plastic layer when the tensile stress is aa in


the bar, so that

c = (1 + cos 20) ,
^ 1 I
(19-7)
oblique now uircn
t = sin 2/3 .
^ Fio. 19-27.
YIELD POINT OF MILD STEEL 321

If the metal in the adjoining thin layer starts to yield and permanent

it,
strains are produced in we may as first approximation neglect the

a
elastic portions of the strains and suppose that the metal yields as under
state of uniaxial simple tension. Let «0 be the small permanent normal
a

strain in the direction of the axial tension stress aa. The metal will
simultaneously contract in all lateral directions perpendicular to the bar
axis and the strains in these directions must be equal to — «„/2. Recalling
from the general equations expressing the distribution of strain for a
permanent uniaxial extension ta in incompressible material that the
normal strain and the unit shear in an oblique direction must be

0
t

equal to

= (1 - cos 20)

3
«

J
(19-7o)
!

}
t = sin 20
^

we may determine the angle of that direction for which the normal
0

strain vanishes. We find


t

cos 20 = V3
= 35°16' . (19-76)
0
,

the weakest material elements in flat bar which are just ready to
If

yield should lie along a line of this inclination, they may do without an

it
additional constraint exerted by the adjoining portions of the bar which
are still in an elastic state of strain. Thus we see that, since the normal
strain in the axial direction has positive value = e„ and the strains
a
«

in the lateral directions perpendicular to the bar axis have negative values
= — ta/2, an oblique direction must exist on the flat sides of the bar
t

along which must vanish and the first small permanent contraction may
e

develop along narrow band without much constraint at the angle


a

= 35° 16'. The first flow layer should then appear as shallow groove
a
0

on both flat sides of the bar at this inclination. If the bar has weaker
a

region in its interior, two slip layers may form symmetrically inclined
with respect to the axis intersecting each other in the weaker region.
A

good example of similarly inclined parallel flow layers was described


recently by A. H. Stang, M. Greenspan, and S. B. Newman,1 photo
a

graphic view of which reproduced in Fig. It shows flat bar


is

19-28.
a

of aluminum alloy 24ST tested by them in tension on which the slip lines
became visible after the scale on the flat sides of the test bar had been
removed by emery paper after the test was interrupted. This alloy
had a well-defined yield point. The parallel and regular spacing of these
Poisson's Ratio of Some Structural Alloys for Large Strains, J. Natl. Bur.
>

Research
Standards, vol. 37, paper RP1742, October, 1946.
322 THEORY OF FLOW AND FRACTURE OF SOLIDS

straight slip bands is most remarkable. Their angle of slip was approxi
mately 57 deg. The direction of rolling was parallel to the bar axis.
The authors report that these equidistant slip lines appeared succes
sively, one at a time, accompanied by small abrupt drops of the load and
by audible clicks and after the specimens had
been stretched 2 to 3 per cent. Strain aging
may have progressed simultaneously during
their formation.
The Dutch engineer P. P. Bijlaard1 de
serves credit for having further analyzed the
various mechanical circumstances which
seem to contribute to the development of the
oblique fracture of wide flat bars. The
following remarks were influenced by
Bijlaard's paper quoted below but are pre
sented in two alternative ways.

Suppose now that a flat bar of ductile metal after


having been extended permanently to the maximum
load is just ready to neck down and that a\,<xi,<ii
denote the principal stresses in an element of metal
in this instant when the lateral thinning along a
straight line inclined at an angle (} (we may utilize
Fig. 19-27 again) starts. From Eqs. (19-7), it fol
lows that

tan 20 = (19-S)
2a

Fig. 19-28. Parallel flow layers We may assume that the principal stresses after
produced on flat bar of alumi
having been raised to the values ffi,o«,<Tj satisfy a
num alloy 24-ST after inter
mittent loading. (Acknowledg condition of plasticity
ments are due to A. H. Stang,
M. Greenspan, and S. B. New (<T, - <72)2 + W, - 03)' + (ffj
- <T,)» = 2*<? (19-9)
man, Bureau of Standards,
Washington, D.C.) where <r<>is the yield stress in tension in the strain
hardened condition of the metal just before neck-
ing sets in. Assuming a state of plane stress in the oblique layer the principal stress
in the direction perpendicular to the flat sides of the bar is o» =0; therefore, from
Eq. (19-9)

0 1(72 + ff!S = (To' (1»-10)

Using the components <rt,o,,T„ of this state of stress referred to the axes z and y

1 "Theory of Local Plastic Deformations in Structural Steel," Institute of Tech


nology, Bandoeng, Java (De Ingenieur in Ned. Indie, No. 8, p. 129, 1940). Also Rept.
Intern. Assoc. Bridge and Structural Engrg., vol. 6, p. 27, Zurich, 1940-1941.
YIELD POINT OF MILD STEEL 323

shown in Fig. 19-27, we have

<n —
— -T—
"J »»
s
.
~r rmu
2
<Tr + «V
(19-11)
— , .
<r»)'
V(<r* 4
r Tx,' .

After inserting these values in Eq. (19-10) the plasticity condition is also expressed by

«** - OxOy + ff»* + 3rI1(S


= oj . (19-12)

The small permanent normal strains «,,«, must be equal to [see Eqs. (16-14)]

(19-13)

The strain <„ = 0, however, in the direction parallel to the edges of the oblique layer
when it starts to neck down because the adjoining portions of nonyielding material
do not participate in the deformation.' Hence

'.-^ (19-14)

and from Eq. (19-12),

So,* + 12t,„j = W . (19-15)

When the plastic limit is just reached in the thin flow layer we see that the two com
ponents of stress o, and t,„ according to Eq. (19-15) must satisfy the equation of an
ellipse having the major and minor semiaxes equal to 2o0/v'/3 = 1.155a0 and <ra/\rZ
= 0.577*0 .

Since the direction of the principal stress ai in the flow layer is parallel to the axis
of the bar we must have in the plastic zone:

tan 2/3
= -^-
at —
= ^*
ai

(19-16)
av

1 Also, since the normal stress a, vanishes (<rt


= o> = 0) and the strain

* "
2~
{"■


)

a strain «, will develop in the directions normal to the flat sides of the bar in the
plasticized layer equal to
" ~ <ft(gj + "y) _" — 3<frTi
''
4
2

expressing the strains with which the layer contracts corresponding to which shallow
a

groove will appear on each flat side of the bar.


324 THEORY OF FLOW AND FRACTURE OF SOLIDS

Taking in Kqs. (19-7) and (19-8) for a = aT and for t = t,„, and after equating the
right sides of the two Eqs. (19-8) and (19-18) expressing tan 2/3, we see that

2ov,
(19-17)
3

From the first of Eqs. (19-7) it follows that

|3
= 35°16' (19-18)

and from the second of Eqs. (19-7) that

Also: (19-19)
(ff, — <7-,)»
+ T»»

As the condition of plasticity [Eq. (19-15)] finally shows the tensile stress a„ in the bar
under which the flow layer forms
<7„ = <ro (19-20)

becomes just equal to the yield stress in tension o0.


We see that, in contrast to the mode of yielding described in Sec. 19-3 for compact
specimens through the formation of a flow layer inclined at an angle 0 = 45 deg in
which a state of plane strain prevails and which will form when the (upper) yield
stress reaches the value aa = 1.155oo, <ro being the lower yield stress in tension, wide
flat bars may start to yield in an oblique line inclined at an angle of 0 = 35°16' under
a tensile stress <r„ = <ra, and a necked portion will form by thinning along this line
under the principal stresses:

On ■ P"q
<Tl
2 ~*"2
(19-21)
a-i
2 2
<ri = 0.
Excluding strain aging phenomena, for the formation of the first flow layer or pair
of crossed layers in a flat bar a state of uniaxial simple tension suffices, while the
initiation of the yielding in a thin layer traversing a massive bar is considerably
retarded through the constraint of the adjoining elastic portions of the specimen.
We may compute the critical value of the angle (3 by a third method as follows.
Equation (19-10) in the rectangular variables <ri and ai expresses the equation of a
"plasticity ellipse" [see Eq. (15-21)] which is the intersection of the general surface of
yielding f(a,,<ri,at) = 0 [Eq. (19-9)] with the coordinate plane <ri =0. By a change of
the coordinates:
ri + <rt = V 2 at' ,

T\

ff2 —

V 2 <Tt , (19-22)

Eq. (19-10) is transformed into

(*,')' + 3(»,')' = 2<r„» (19-23)

showing that the plasticity ellipse has the semiaxes \/2 <r»and *3 O"0.

*
YIELD POINT OF MILD STEEL 325

Any state of stress in the flow layer is represented by a Mohr circle in the rec
tangular variables a (normal) and t (shear stress) :

(gi:g,)'- (19-24)

In this equation <n and at for those states of stress which just reach the plastic limit
must satisfy the plasticity condition Eqs. (19-10) or (19-23). All Mohr circles repre
senting the latter states of stress in the flow layer have an enveloping curve whose
equation will be found by eliminating from Eqs. (19-10) and (19-24), for example, oi
and by considering n as a variable parameter in the resulting equation. Instead of
proceeding in this way, we may make use of the coordinate transformations [Eq.
(19-22)] and after denoting by p the variable parameter

express Eq. (19-24) also as follows:

(a
- p)» + t» =
IW
- p>) . (19-26)

This last equation represents family of Mohr circles F(<r,T,p) = 0 in


a one parametric
the variables <rand t which has an envelope whose equation is found by eliminating p
from this last equation and from the condition dF /dp — 0, which gives for p = 3a/4.
For the envelope of the stress circles the equation of an ellipse

&»» + 12r« = W (19-27)

is obtained which coincides with our former Eq. (19-15). Among all states of stress
at the plastic limit, represented by their Mohr circles [Eq. (19-24) or (19-26)], i.e.,
whose circles must be tangent to the ellipse Eq. (19-27) for material elements situated
in the flow layer according to Bijlaard1 we wish now to determine the state of stress
for which (1) the condition «„ = 0 holds, i.e., for which according to Eq. (19-14)
<rt
~ <r«/2 and (2) the tensile stress o„ in the flat bar will become a minimum. The
resultant stress o„ in plane AB (Fig. 19-27) is

<Tr„2 = a' + T* . (19-28)


Since a = az = o>«, cos 0 , 1
r = tiv = a„.sm p ) (19-29)

after substituting these values in Eq. (19-27), we see that

2<T„

V3 Vl + 3 sin» 0
(19-30)

and that the tensile stress <ra is equal to

_ jr™ m 2^
1

cos0 V3 cos 3 Vl + 3 sin2 ^

1 hoc. cit.
326 THEORY OF FLOW AND FRACTURE OF SOLIDS

The stress <r„ becomes a minimum for daa/d0 = 0, which condition determines the
unknown angle 0 and gives again
cos/3 = V%, 0 = 35°16' (19-32)

in agreement with our former computations. With this value of /3 we obtain from
(19-31) the tensile stress

<r„ = <r0. (19-33)

If an angle /3 of the flow layer is as


sumed, for example, equal to zero, from
Eq. (19-31) <r„ = 2<ro/v/3 = 1.155<r0 and
if 0 is chosen equal to /3 = 45 deg,

<r« = 1.033<To

is obtained, in both cases larger stresses a,


than an = <r0. Thus we note that in flat
a..a. —- bars of isotropic metal the local distor
tion develops at 0 = 35°16', because this
Fio. 19-29. Ellipee: 3a' + 12rJ 4<ro
requires the least amount of energy of plas
tic distortion or the smallest total tensile load (stress a,).'
The Mohr circle satisfying the preceding condition
for a minimum of the tensile stress ao and tangent to
the ellipse Eq. (19-27) is shown in Fig. 19-29, from
which it is noted that the line connecting the origin O
with point P represents the resultant stress <rre, and
the angle under which OP is inclined with respect to
the major ellipse axis OA is the angle 0 = 35°16'
which has just been computed.
The fact that a, as a function of the angle of slip 0
according to Eq. (19-31) has a comparatively flat
minimum perhaps explains the variations in the values
of the angle of slip which have been observed in the
experiments. After applying two equal small but
sufficiently deep and sharp notches opposite each
other on the narrow edges of a flat specimen, a flow
layer at an angle /3 = 0 may form more easily as the
oblique layer, as was verified in old experiments by
the author.
That conditions may become more complex is
illustrated in the photograph of a flat steel bar (Fig.
19-30) taken just before it broke. The ratio of Fio. 19-30. Flat teal bar
tested in tension n which
original width to thickness in this case was 10, just
complex necking developed.
about of sufficient magnitude to cause symmetrical Original dimensions of bar:
necking in combination with the beginning of thin width 3.75 in., thickness
ning along two oblique directions. The reflections of 0.375 in., gauge length 18.75
in.
light on the flat side of the specimen clearly disclose
1 As Bijlaari) clearly points out in a second paper
(On the Restricted Applicability
of the Principle of Least Work in the Plastic Domain, Proc. Koninkl. Nederland.
Akad. Wetenschap., vol. 1, No. 4, 1947, Amsterdam), the bar adjusts its mode of flow
YIELD POINT OF MILD STEEL 327

a white cross with oblique angles, indicating the regions along which a shallow
depressionon the flat sides developed before fracture occurred.1 Figures 19-31 and
19-32 show symmetrical necking in flat bars when the ratio of width to thickness

Fio. 19-31. Fio. 19-32.


Figs. 19-31 and 19-32. Symmetric necking in flat steel bars, h = 0.75 in. b/h =7.
h = 0.187 in. (Teats by J. Miklowitz.)
was equal to 7.2 The theory of the oblique fracture of wide flat bars has been
extended to anisotropic flow in strain hardened metals by R. Hill.3

in accordance with the requirement of a minimum amount of energy of permanent


distortion, provided that the small amount of simultaneously stored elastic energy is
neglected. In other words, a minimum of the energy occurs if only the permanent
parts of the strains are included. Regarding the variational principle of the least
amount of energy occurring under a plastic distortion, the reader is referred to Vol.
2. The state of stress determined in the text above does not correspond to a mini
mum of the total (elastic plus plastic) strain energy, as one might perhaps be inclined
to assume.
1 This bar still broke in symmetrical way in a shear fracture; the traces of the shear

crack on the flat sides were at a direction perpendicular to the bar axis.
2 The influence of geometry
(ratio of width to thickness) and of size was carefully
investigated for flat bars broken in tension tests by J. Miklowitz in a paper presented
at the ASME Symposium of Flow and Fracture of Metals, June, 1948, Chicago, 111.
(J. Applied Mechanics.)
3 Loc. cit.
CHAPTER 20

COMPRESSION
Although no difficulties are encountered in obtaining a uniform stress
distribution in a bar of ductile material under tension until necking
begins, special precautions are necessary in order to load a cylindrical or
prismatical bar under axial compression between two hard plates in such
a manner that the lateral extension of all cross sections is the same. Since
that part of the test piece which touches the plates producing the com
pression is hindered from expanding laterally because of the friction, a
bulging in the middle part occurs and there results, especially in the
neighborhood of the ends of the test piece, a nonuniform distribution of
compressive stress.

Fig. 20-1. Compression test. Flow figures (Li'iders' lines) on the four faces of a prism
of mild steel. (Photographs of highly polished surface of specimen reproduced in Figs. 20-1
and 20-2 were made by using the Schlieren method described in Sec. 18-1.)

The development of the flow layers in the case of compressed mild steel
specimens, shortly after the yield point has been passed, is illustrated in
Figs. 20-1 and 20-2; also Figs. 18-37 to 18-44.
cf.

20-1. Cylinders. a compression test made in testing machine


If

is

with a cylindrical specimen of brittle material, such as sandstone,


a

marble, or concrete, or of a ductile metal, certain more or less regular


markings will be noticed on the specimen before failure. The main thing
which will be observed that the greatest distortion of the material in
is

the specimen usually begins at the edges of the compression surfaces and
generally more or less concentrated along two conical surfaces starting
is

from the ends of the cylinder. In this way two pieces of conical shape
328
COMPRESSION 329

split off from the rest of the specimen before failure in the brittle case of
materials, and under a larger load these pieces become wedged into the
middle part of the specimen and cause this to crack (cf. the compressed
sandstone cylinder, Fig. 20-3).

(a) (6)
Fig. 20-2. Flow figures on mild steel. Compression tests.

On well-polished cylinders of white marble during a compression test


two very regular systems of helical lines may be seen on the surface.
Both these systems of lines make the same angle with the direction of

Fio. 20-3. Sandstone cylinder broken in compression test exhibiting cone of fracture.
(According to tests of T. von Kdrmdn.)

compression. They are the traces on the surface of the cylinder of two
series of helicoidal surfaces in which the marble is more severely distorted
than in the neighboring parts. It is possible to make these severely
distorted layers visible to the eyeif the surface is colored or rubbed by
means of a colored pencil. In this way, the markings on the surface of
330 THEORY OF FLOW AND FRACTURE OF SOLIDS

an axially compressed cylinder of Carrara marble (Fig. 20-4) were


obtained. (This cylinder was part of a series of tests on various kinds of
rock by Prandtl and Rinne.) A second example of the markings on the
surface of a compressed-rock cylinder is shown by the photograph of
Fig. 20-5. These regular markings indicate that in these tests a distribu
tion of stress with axial symmetry under compression existed.

Fig. 20-4. Helical slip lines on polished marble


cylinder after compression. (Test by Prandtl-
Rinne.)

Fio. 20-5. Helical slip lines on stone cylinder


after compression test.

In order to investigate these phenomena in plastic materials the author


made a series of compression tests using very soft materials.1 These
tests will be briefly reviewed in the following:

A dark-colored paraffin having the property of becoming light wherever excessive


distortion of the material occurred was used. By means of the resulting bright lines
and markings obtained on the surfaces and cross sections of these stressed paraffin
specimens, the regions in which a severe distortion of the material existed were made
plainly visible.
After the maximum load on these compressed paraffin cylinders is reached, mark
ings consisting of fine lines arranged in a more or less regular way develop. These
lines, which are similar to those in the marble specimens, consist of two systems of
helical lines crossing each other. These lines make angles of about 45 deg (usually a
little less than 45 deg) with the generatrices of the cylinder. During the course of
the test, single lines or groups of lines become more pronounced until fracture occurs

Cf. Z. tech. Phys., vol. 5, No. 9, p. 369, 1924; and Proc. Intern. Cong. Applied
1

Mechanics, p. 318, Delft, 1924.


COMPRESSION 331

Fio. 20-6. Helical systems of slip lines on surface of paraffin cylinders tested in compression

Fio. 20-7. Fio. 20-R.


Figs. 20-7 and 20-8. Helical systems of slip lines on surface of paraffin cylinders tested in
compression.
332 THEORY OF FLOW AND FRACTURE OF SOLIDS

with a resulting drop in the load. In Figs. 20-6 to 20-8, the markings obtained in a
series of cylindrical test specimens of five various heights are shown.1
In Figs. 20-9, 20-12, and 20-14 a few longitudinal cross sections of compressed paraf
fin specimens are represented. These sections were taken through the axes of the
specimens. Figure 20-13 represents a schematic sketch of the structural changes

Fig. 20-9. Fio. 20-10.


Figs. 20-9 and 20-10. Structural changes in longitudinal cross section (left) and on surface
(right) of a paraffin cylinder tested in compression.

observed in the cross sections of the three cylinders shown in Fig. 20-6. An external
view and a longitudinal cross section of a truncated cone, having a small angle of cone
and tested rapidly in compression, are shown in Figs. 20-9 and 20-10.
It will be recognized from the longitudinal cross section (Fig. 20-9) that the curves
of constant degree of destruction of the material (and of constant intensity of color)
have approximately the shape shown by the lines of the
sketch of Fig. 20-11. A more careful consideration of such
cross sections shows that the amount of distortion of the
material inside the specimens (the more the material has
been deformed the brighter it appears in the photographs)
changes throughout the inside of the specimen in a regular
Fio. 20-11. Section
through axis of a manner. Two truncated cones of material at each end of
cylinder of soft material the specimen seemed to adhere to the plates of the testing
near compression plate, machine used to transmit the compressive load. (In order
showing areas of equal
to produce a strong adhesion the paraffin specimens were
degree of destruction.
compressed between steel disks having the same radius and

1 At this point may bo mentioned a remarkable phenomenon which was brought out
by a series of slowly made compression tests and a second series of rapidly made tests.
In the last case, the loading was brought upon the test piece very suddenly, such as
would occur if the heavy pendulum of an Amsler testing machine were lifted high and
then suddenly released, while the test piece was under a small compressive load.
During the rapidly made tests the surface of the cylinder was covered with a much
finer and denser network of helical lines than was the case with the slowly made tests.
COMPRESSION 333

intentionally rough-machined surfaces.) These two cones, for example, can be seen
in the dark part of the longitudinal cross section of a specimen shown in Fig. 20-12.
In the cross sections the compressed regions a (Fig. 20-13) have a black and white1
mottled appearance with a fibrous structure whose direction was perpendicular to

Fio. 20-12. Section through axis of a paraffin cylinder after severe compression. If the
cylinder had been stressed more highly, the cones of fracture would have developed from
the four narrow bright sectors originating in the corners.

~5|

Fig. 20-13. Fio. 20-14.


Fio. 20-13.Longitudinal section through three cylinders of different heights from a soft
material, showing structural changes due to compression test.
Fio. 20-14. Section through a cylinder of paraffin with constricted portion exhibiting two
cones along which fracture would have developed under further loading.

the direction of compression. If the two cones extended uniformly into the test piece
this mottled surface extended quite deeply toward the middle of the test specimen.
1 The
plain white surfaces (not crosshatched) in the sketches (with exception of the
cones 6) likewise represent regions of changed and distorted texture, in which, however,
no definite fibrous structure was recognizable (see Fig. 20-12).
334 THEORY OF FLOW AND FRACTURE OF SOLIDS

This, however, was usually the case only for the shorter specimens. Areas of dis
tortion c (Fig. 20-13) extended like rays from the edges of the cylinders, both systems
of slip lines being often very noticeable in these areas. (For example, compare the
section through a severely distorted cylinder, Fig. 20-12.) No slip lines were visible
along the axis of the cylinder, a fact which is required from conditions of symmetry.

Fio. 20-15. Fio. 20-16.


Figs. 20-15 and 20-16. Slip lines in section of rings loaded in axial direction by com
pression.

The angle which the ray-shaped areas of distortion make with the axis of the speci
men depends, for the same material, on the friction existing in the compressed surfaces
of the specimen and on the ratio of length of cylinder to diameter. Using similar
compression plates, three cylinders, having different values of the ratio h/d of height

Fig.20-17. Fig. 20-18.


Figs. 20-17 and 20-18. Section of two paraffin rings through axis exhibiting structural
changes due to compression.

of cylinder to diameter, gave the following values of the angle a between the gener
atrices of the cone of rupture and the base:

Value of h/d Value of a


1.00 36°
0 50 26°
0 33 2,o

A few of the shorter cylinders occasionally showed on their surface, besides the
helical slip lines, countless fissures, running parallel to the direction of compression.
These tended to open up under further loading and frequently, under close examina
tion, showed a tendency to form jagged lines because of crossing the helical slip lines
COMPRESSION 335

The deformation of a few rings, loaded axially in compression between smooth hard
plates, will now be considered. The deformation of two rings is represented in an
easily understandable way by the slip lines in Figs. 20-15 and 20-16. In the case of a
thin ring (Figs. 20-15 and 20-17) the deformation was quite different from that in the
case of a thick ring (Figs. 20-16 and 20-18). Since the conical slip surfaces in a thick
ring come to an intersection at the inner surface of the rings near the middle of the
height, the displacements were such as to form a ring-shaped swelling on the inside
surface of the ring. This bulged-out portion is easily recognizable in the photograph

Fio. 20-19.

Fio. 20-20.
Figs. 20-19 and 20-20. Plane compression tests. Sections through the middle plane of
two flat prisms of paraffin subjected to compression in a direction parallel to small edge of
rectangle. The distortion of the section can be recognized by the regular markings which
were formerly circular.

of the longitudinal section. No such deformation could be observed with the thin
rings.
The distortion was axially symmetrical only for the case of the shorter cylinders;
as soon as the height of the cylinder exceeded a certain value, the distortion often
became unsymmetrical and was, moreover, limited to a short portion of the length
(c/. Figs. 20-7 and 20-8). Cylindrical specimens of brittle material often fracture so
that the final surface is composed of portions of both a right- and a left-handed heli-
coidal slip surface together with a part of the cone of slip extending from the com
pression plates. These three surfaces unite to a common surface of fracture, which
sometimes may not differ much from an oblique plane.

20-2. Plane Compression Tests. Prisms having narrow rectangular


cross sections in planes perpendicular to the direction of compression
were tested between two compression plates. These gave slip lines such
as those shown in the photographs of Figs. 20-19 and 20-20. Before the
336 THEORY OF FLOW AND FRACTURE OF SOLIDS

tests, the prisms were cut through along a middle plane coinciding with
the direction of compression and during the tests both pieces were held
together by light pressure. The mode of deformation of the prism is
rendered visible by noting the distortion of the grooves left by the cutting
tool of a lathe. The test specimens, whose middle cross sections are
shown in Figs. 20-19 and 20-20, had approximately equal widths (33 to
35 mm) and lengths (125 mm) but a different ratio of height to width.
When the deformation had proceeded sufficiently and had occurred
symmetrically, certain regular systems of slip lines extended from the
corners of the test piece under compression. These crossed each other
in the middle portion of the test specimen in a
manner described by L. Prandtl. ' It will be noted
that the originally concentric circular tool marks
become ellipses only in the central portion of the
Fio. 20-21. Sketch in- cross section. In the four outer areas of triangular
dicating cross section of
a flat prism before (left
half of figure) and after
... ,. ....
shape (see Fig. 20-19) the circles remain almost
..
concentric, while the material in four narrow sectors
(right half) a plane
compression test.
rection of compression
Di-
emanating from the corners must have been sub-
, ,. .. 1.1 i • i-i
jected to extraordinarily high shearing displace-
stress parallel to short ments. As a consequence of this nonuniform defor
mation the material is distorted mainly in these
narrow sectors so that finally the test specimen breaks into four pieces.
This is indicated by the schematic sketch in Fig. 20-21.
20-3. Prisms. A series of prisms with square cross sections of 36 mm
length and having heights of 60, 48, 36, and 24 mm, when tested ga%e
the slip figures shown in Figs. 20-22 to 20-28. The prisms were com
pressed between wood plates having rough-grooved surfaces. The rela
tionship between these slip figures and those of the plane problem is
unmistakable. However, in this case, the converging systems of slip
lines make a greater angle with the direction of compression, correspond
ing to the different (radially directed) direction of the friction stresses
in the compression planes. In the case of the shorter prisms, four por
tions of the specimen along the four edges parallel to the direction of
compression remained undeformed, while the middle portion of the faces
of the prism showed a network of slip lines of marked regularity. A
bulging out of the middle portions of the sides was also noticeable. If
we think of a cylinder inscribed in the prism, its axis coinciding with that
of the prism, and the portions of the material near the corners cut away
in the manner shown in Fig. 20-29, we shall recognize the network of
slip lines of the prisms as a part of the network of helical lines already

1 Z. Math. Afechanik, 1923.


f. angew.
COMPRESSION 337

Fio. 20-25.
Fios.20-22 to 2,-25. Compression tests with paraffin prisma of same square cross section
and of different heights.
338 THEORY OF FLOW AND FRACTURE OF SOLIDS

Fig. 20-26.

Fig. 20-27.

Fig. 20-28.
Figs. 20-20 to 20-28. Systems of slip lines appearing on sides of paraffin prisms of square
cross section tested in compression. The effect of the height of the prisms on the shape of
the slip lines is noticeable.
COMPRESSION 339

described for the case of cylinders under compression. They appear on


the surface of the prisms, where the inscribed cylinder touches it.
Similar slip lines are exhibited on the surface of compressed marble
specimens, as, for example, on the photograph of Fig. 20-30, showing one
of a group of specimens of Carrara marble, tested by
7^ Prandtl. Moreover, in this case, the remarkably steep
angle of the marble slip lines is exhibited.
Fig. 20-29. 20-4. Compression Tests on Cylinders of Ductile Metals.
Although the above described compression tests were carried
out only with a material having a very loose texture or on brittle rocks,
the phenomena observed will apply at least qualitatively also to com
pression tests of ductile metals. This has been established by countless
observations1 and tests. The action of the friction between the com
pressed plates is exhibited by the well-
known barrel-shaped form which the
shorter metal cylinders assume under com
pression. Where the cylinders have
lengths from one and one-half to three
times the diameter and where ductile
metal is used, the deformation is usually
such that two symmetrical enlargements
or a bulging out of the cylinder in two sec
tions is observed. For longer cylinders,
sidewise buckling may occur.
Regions of large shearing deformations
diverging from the edges of the cylinders
may also be shown to exist in longitudinal
sections of cylinders of mild steel. The
lines or markings which result in the
etched longitudinal section of the test
specimen, because of the presence of the
inclusions and segregated zones in the
Fia. 20-30. Slip lines and surfaces
ingot, permit one to determine how a of fracture of a marble prism after
mild steel cylinder deforms when com compression test. One of the
" pyramids of fracture " can be seen
pressed between two hard-steel plates. in the upper part of the test piece.
In order to decrease the friction between
the compressed surfaces, Hiibers2 lubricated the compression plates with
1
Cf. the compression tests of Riedel, Hiibers, Meyer and Nehl, Siebel, etc., men
tioned below.
* Das Verhalten einiger technischer Eisenarten
beim Druckversuch, Wahwerkaus-
schussberichi des Verein deulscher Eisenhilttenleute, No. 32, Diisseldorf. This paper
gives considerable data relative to compression tests.
340 THEORY OF FLOW AND FRACTURE OF SOLIDS

a mixture of oil and graphite. He thus obtained a series of markings as


shown in Figs. 20-32 to 20-36, which are copied from his photographs.
While the cylinders lubricated with graphite expanded laterally in a uniform
manner, the unlubricated test specimens showed a barrel-shaped form with

i
i;

m1
P!

1,1
i

1.
.

;
1
Fio. 2,-31. Fio. 20-32. Fio. 20-33.

F
Ell
111 111
Y

Fio. 20-34. Fio. 20-35.


Fio. 20-31. Section through steel cylinder after was pressed between hard plates,

it
a

showing the distortion of the parallel sets of streaks due to the impurities of the steel.
(According to H. Meyer and F. Nehl.)
Flos. 20-32 to 2,-35. Longitudinal sections of steel cylinders after compression tests.
Left figures: lubricated with graphite in compression plane. Right figures: without
lubricant. (According to K. Hubero.)

considerable distortion of the system of longitudinal lines or fibers pro


a

duced by rolling. (Comparing the markings of Figs. 20-13, 20-31, and


20-35, should be noted that the conical surfaces exhibiting large shear
it

ing deformations in the highly compressed steel cylinders are continually


forming anew, and their position changes relative to the test specimen.
The cylindrical surface near the ends of the cylin
Orifinelly o
y

der tends to invert, and gradually moves so that

it
part at the
Cylindrical
Surtou becomes part of the outermost ring-shaped part x
of the circular area under compression. This

is
shown by section of a specimen in Fig. 20-36.)
a

Meyer and Nehl1 have further investigated the


Fio. 20-36. Base of a phenomena in compressed iron cylinders by means
test cylinder after severe of etching the longitudinal sections, using Fry's
compression. Distortion
of the formerly parallel method. They have also investigated the change
streaks due to the im in load under compression. In the schematic
purities, the orientation
of which was here per reproduction of an etched longitudinal cross section
pendicular to axis of of compressed iron cylinder (Fig. 20-37) the con
a

cylinder.
ical areas in which the friction of the compression
plates hinders the lateral expansion may be clearly recognized. In test
a

specimen with very accurately machined ends Meyer and Nehl were able
to demonstrate clearly the presence of the conical surfaces of slip (appear-
Die grundlegenden Vorgiinge der bildsamen Verformung, Stahl u. Eisen, vol. 45,
1

p. 1961, 1925 (Mitt. Prufungsamtalt der August Thyssen-HiUte, Hamborn).


COMPRESSIOX 341

ing asblack lines in Fig. 20-38). They obtained the following values
for the angle which this conical surface made with the direction of com-

<f>
flfastic

undetormtd
Fig.
20-37. Flo. 20-38.
Figs. 20-37 and 20-38. Region plastically deformed (left) and conical surfaces of slip
(right) in sections of compressed steel cylinders. (According to Meyer and Nehl.)

pression, various ratios of cylinder diameter to height of cylinder being

h
considered:

d/h 0.4 0.6 0.8 1.0 1.2


32° 40" ■17° 55° (52°

The conditions of the ordinary compression test have been substan


tially improved by Siebel and Pomp,1 who suggested that be desired

it
if
to obtain uniform distribution of stress, the cylindrical
a

test specimens should be compressed, not between two


plane plates, but between two cones (Fig. 20-39). If the
generatrices of these cones make an angle with the plane
of compression, equal to the angle of friction, the resultant
stress in the compression surfaces of the two cones
is

Fio. 20-39.
parallel to the direction of compression. This produces Cylinder com
approximately in the test specimen pure axial compres pressed be
a

tween conical
sive stress. end plates.
(According
For an ideal compression test with uniform and unre to
E. Siebel.)
stricted lateral expansion, we have, as in tensile tests, since
the volume of the compressed cylinder practically constant
is

■Kr*h = xro^o (20-1)


,

where and are the radius and height, respectively, of the cylinder at
h
r

any given time, and r0 and ho are initial values of and The trajectory,
h.
r

which an arbitrary point describes during an ideal compression test,


r,
h

Die Ermittlung der Formandcrungsfestigkeit von Metallen durch den Stauch-


1

versuch, Mitt. Kaiser-Wilhelm Inst. Eisenforsch., vol. p. 157, Diisseldorf, 1927.


9,

Cf.
also E. Siebel, "Grundlagen zur Berechnuug des Kraft- und Arbeitsbedarfs beira
Schmieden und Walzen," Dr.-Ing. Dissertation, Berlin, 1923.

r
342 THEORY OF FLOW AND FRACTURE OF SOLIDS

is given by the equation

h =
^ •
(20-2)

Defining the unit compression in axial direction as a positive quantity


by putting it equal to

e =
^f-* (20-3)

and similarly by denoting the compressive (true) stress taken with respect
to the actual cross section xr2 by its absolute value a, the work done in
uniformly compressing a cylinder from a height h0 to a height h is found
equal to

W = -x Jh,
[ rV dh = -TToVio
[
Jh,
* ?
n

(20-4)

If in this expression we substitute for h the unit compression e, we


obtain [since from Eq. (20-3) h = h0(l — t), dh = — h0dt] the work of
deformation equal to

ir = xro^o
/ •
(20-5)
Jo 1 — «

xr02Ao is the volume of the test cylinder. Thus the integral represents
the work done in compression per unit volume, which we may denote
by w. In Eq. (20-5) we may make use of the natural compression strain i
in axial direction, which we may define in a manner analogous to the way
the natural strain in tension was introduced in Eq. (8-7) by letting

~<=
- Ytt =
lnTn = ~ln (l ~ «)• (2°-6)
iuj n
Thence

i -
dt
«

If we consider the stress a in terms not of « but of the natural unit com
pression i
a = F(i) (20-8)

we obtain the work of deformation per unit volume in the simple form

w = Y <rdl. (20-9)

This is the area under the natural stress-strain curve a = F(i) for com
pression.
In Figs. 20-40 to 20-45, the deformation curves in tension a = /(«)
and compression for soft iron, two carbon steels, a nickel steel, copper,
COMPRESSION 343

and aluminum, by Siebel and Pomp, are given.


as determined In all
cases the curves of stresses in the plastic stage of iron and steel for com
pression are higher than the corresponding ones for tension. For copper

lb$./in?
150,000
ibi./in.2

0 20 40 60 80% 20 40 60 80°/. 20 40 60 80%


e e
Fig. 20-40. Wrought iron, Fig. 20-41. Carbon steel, Via. 20-42. Carbon steel,
0.06% C, 0.15% Mn. 0.21% C, 0.17% Si, 0.62% 0.39% C, 0.72% Mn.
Mn.

Ibs./t'n.1
190,000
Ibs./in?
150,000 - 15,000

lbs./inf
45,000

0 20 40 60% 20 40 60% 0 20 40 60 80°/.


C t
Fig. 20-43. Nickel steel, Fio. 20-44. Copper. Fig. 20^45. Aluminum.
0.23% C, 0.65% Mn,
24.6% Ni.
Figs. 20-40 to 20-45. Deformation curves a — /(e) for tension and compression. (Accord
ing to Siebel and Pomp.) (Abscissas: unit elongation or compression «. Ordinates: true
tensile or compressive stresses a referred to actual cross-sectional area.)

and aluminum the a curves for compression differ but little from those for
tension, in agreement with earlier tests of Ludwik and Scheu.1
20-6. Hollow Cylinders of Porcelain in Compression. The shape of the heads of
hollow cylinders of a brittle material stressed in axial direction by a compression load
and the conditions in the contact area between the cylinder and the hard steel plates
have a great influence on the compression strength which might be observed. This is
illustrated in the results of tests made by J. S. Lambert and M. J. Manjoine2 with hol-
1 Vergleichende Zug-, Druck-, Dreh-, und Walzversuche, Stakl u. Eisen, vol. 45,
p. 373, 1925.
' The strength values quoted above are taken from an unpublished report of the
Westinghouse Research Laboratories, May, 1941. Heavily loaded hollow cylinders
of fired porcelain in large dimensions are used to support radio antenna towers. In
such applications difficulties have been observed through cracking in the hollow insula
tors supporting the weight of the steel towers.
344 THEORY OF FLOW AND FRACTURE OF SOLIDS

low cylinders of an insulating porcelain with carefully ground conical contact surfaces
the shapes of which are reproduced in Figs. 20-46 to 20-48. The cylinders were fired
in a kiln. Their outer and inner surfaces were also ground in those cylinders which
were tested in the unglazed condition.
The average values of the observed frac
ture stresses are listed below. The ten
sile strength of the porcelain determined
in bending tests with round bars of * ?f 6
in. diameter was 11,200 lb/in.1 for the
unglazed and 12,800 to 14,000 lb/in. ' for
the glazed bars. (The modulus of elas
ticity in compression was 10,170,000
lb/in.'; Poisson's ratio 0.23.) While
small, solid, round, unglazed and glazed
cylinders (1% in. diameter, 3 in. high)
with ground plane ends broke in com
pression at 73,000 and 83,000 lb/in.*,
respectively, it is most remarkable that
a series of hollow cylinders (the shapes c
and d of Fig. 20-46) whose forms were
suggested by E. E. Arnold with a concave
curved outside surface and having ends
ground according to a cone inclined at an
angle of 12 deg, when compressed be
tween hard steel plates perfectly fitting
Fio. 20-46. Hollow compression cylinders
of fired porcelain. into these cones, sustained compression
stresses (referred to their minimum sec-
tions) as high as 130,000 lb/in.*.1
From these tests the useful conclusion might be drawn that the best shape of a body
of revolution offering the highest strength in axial compression in a brittle material is

Flo. 20-47. Hollow compression cylinder. Fio. 20-48. Toroid body of brittle material
for high compression strength.

1 The
shapes c and d (Fig. 20-46) were proposed by E. E. Arnold, consulting engineer
of the Westinghouse Electric Corporation in East Pittsburgh.

^
COMPRESSION 345

perhaps a hollow cylinder having a variable wall thickness and the middle surface of ti'hich
is a toroid (see Fig. 20-48.) There is a good cause for the superior strength of this
body. It is easy to see that, in accordance with the theory of thin shells of rotational
symmetry, an outwardly concave middle surface, when the shell is loaded in the axial
direction in compression, induces further compression hoop stresses <rt in the circum
ferential directions. It has been mentioned that compact prismatic bodies of brittle
material in compression tests frequently split along surfaces or planes parallel to the
axis of the specimens because of the wedging action of those portions which stick to
the compression plates. Formation of these lateral tensile stresses accompanying
Strength in Axial Compression of Hollow Porcelain Cylinders
(See Fig. 20-46)

Compression strength,
Shape Surface
lb/in.'

Cemented ends (e) Unglazed 64,600

With reinforced shoulders (a) Unglazed 50,200

Straight hollow cylinder (6) Unglazed 56,300


Glazed on outside 70,000
Glazed on both sides 58,600

Cylinder with curved outside sur Unglazed 121,000*


face (c) 66,000t

Cylinder with curved outside sur Unglazed 130,000»


face (d) 117,000f
Glazed on outside 112,400
Glazed on both sides 102,400

Straight solid cylinder (d = 1 )4 in., Unglazed 73,000


h = 3 in.) (plane ground ends) Glazed 83,000

* Highest stress,
t Average stress.
the axial compression test will be retarded where they may start to develop by the
compression stresses in the circumferential directions at induced in the concavely
curved hollow specimen of Fig. 20-48. ' One of these fractured porcelain specimens
is reproduced in Fig. 20-49. In some earlier compression tests with short, solid,
round cylinders of porcelain having ground plane ends, the highest strength values
were observed when a soft sheet of towel paper was inserted between the specimen and
the hard steel plates. When a very thin lead foil was used instead the compres
sion strength dropped to quite low values (indicating that soft plastic lead is readily
squeezed into the invisible crevices of the specimen along the compression plates, and
by its wedging action splits the specimens longitudinally).
effect of these conically ground ends (Fig. 20-46, shapes 6, c, d), is just the
1 The

opposite of that of the cones proposed by E. Siebel (see Fig. 20-39) which served to
prevent retarding of the lateral spreading of a ductile compression specimen through
the radial friction forces along the compression plates.
346 THEORY OF FLOW AND FRACTURE OF SOLIDS

Large hollow cylinders of fired porcelain which are much used in the electric indus
try in insulating bushings of circuit breakers,
etc., are manufactured by extruding the
tubes of wet clay through a die under the pressure exerted by a piston in a steel cylin-

Fio. 20-49. Hollow porcelain cyl- Fig. 20-50. Surfaces of slip in


inder. Broken by axial compres- clay. Section of heavy-walled cyl-
sion along numerous longitudinal inder that broke when it was dry.
cleavage cracks. The tube had been extruded when
it consisted of wet clay. Note the
curved surfaces of slip produced
through the friction along the die
during the process of extrusion.

der. Through this process of plastic forming, difficulties arc sometimes encountered
in the extrusion of heavy-walled cylinders of wet clay. Figure 20-50 shows an exam
ple of the formation of a regular system of curved surfaces of slip in a tube of wet clay
that had been extruded through a ring-shaped die.
CHAPTER 21

TORSION OF A ROUND BAR. THE STRESS-STRAIN CURVE


IN SHEAR
In a cylindrical bar of circular cross section permanently twisted by
two moments by a small angle around its axis, the distribution of stresses
in a cross section can be determined by simple means for an isotropic
material deforming under an arbitrary law.1 For comparatively small
angles of twist the distortion of the cylinder may be assumed to consist of
simple shearing strains 7 which are proportional to the distance r of a
point P in the bar from the axis. This amounts to assuming that two
plane sections originally perpendicular to the axis a distance I apart will
rotate with respect to each other around the axis through a small angle
which is proportional to I

a = 61 (21-1)

and that these sections do not change their distance. The quantity 6
defines the angle of twist of the bar per unit length. No components of
stress normal to these sections need be considered, and
the same is true for the components of normal stress
in the radial or circumferential directions. The only
components of stress not vanishing will be the shearing
stresses t in the plane sections perpendicular to the axis
whose arrows are normal to the radii r and in the planes
containing the axis of the bar whose arrows are parallel
to the latter.
Considering two neighboring cross sections a distance dl apart, a point
P (Fig. 21-1) at a distance r from the axis is displaced, when twist occurs,
along a small circular arc PP'. Under this displacement an element QP,
which was initially parallel to the axis of the bar, is tilted by the small
angle 7 equal to the infinitesimal unit shear and becomes an element
QP' along a helix very steeply inclined to the planes of the cross sections.
1 This apparently was first suggested by Ch. Duguet, "Limite d'<51asticit(5 et resist
ance a la rupture," vol. 1, p. 157, Paris, 1882; then independently by P. Ludvvik,
"Elemente der technologischen Mechanik," Verlag Julius Springer, Berlin, 1909; and
by L. Phandtl, in his university lectures, cf. Herbert, Mitt, u. Forschungsarb. Ver.
deul. Ing., No. 89, 1910.
347
348 THEORY OF FLOW AND FRACTURE OF SOLIDS

We have
PP' = rda = rddl = y dl (21-2)
defining the unit shear
7 = rd . (21-3)

Suppose now that the unit shear in the material depends on the shear
ing stress t and that t = /(>) is a monotonously increasing function of the
unit shears y,
t = J{y) (21-4)
is called the stress-strain curve of the material for shear.
a. If
this curve is known for a material it is easy to determine the distribu
tion of the shearing stress r in the cross sections of a twisted round bar
under a given twisting moment M and to find the function M = F(0), or
the moment-twist curve, which may easily be observed in a torsion test.
Let ya = ad equal the shear y when r = a, a denoting the radius of the
round bar.

UNIT SHEAR

Fig. 21-2. Shearing stresses in perma Fiq. 21-3. Stress-strain curve in shear.
nently twisted round bar.

On account of the proportion r/a = 7/7,, we


see that the curve repre
senting the shearing stresses t
function of the radii r in a cross section
as a
of the bar (Fig. 21-2) has the same ordinates t as the stress-strain curve
f = f(y) (Fig. 21-3) when the ratios r/a and 7/70 are the same. After
having constructed the curve t as a function of r corresponding to a given
angle of twist 0 = yja or to a given value of the unit shear ya — ad
at the circumference of the boundary circle r = a, the twisting moment M
is given by the integral

M = 2t f" rr* dr (21-5)

or, if instead of the variable r the variable 7 is substituted according to


Eq. (21-3) r = y/0, dr = dy/0, also by

,V ~~
/(7b2 dy . (21-6)
03 Jo
TORSION OF ROUND BAR 349

The twisting moment M is equal to the moment of inertia of the shaded


area under the stress-strain curve for shear t = 7(7) (Fig. 21-3) taken
about the t axis, this multiplied by 2ir/63.
b. Conversely, if the moment-twist curve M = F(6) is known from obser
vation in a torsion test, the unknown stress-strain curve for shear t = f(y)
of the material may be determined from Eq. (21-6). The right side of the
equation

MO3 = 2t fj'f(y)y*dy (21-7)

is a function of the upper limit ya and since ya = a6, it is also a function


of 6. After differentiating (21-7) with re
spect to 6,
, M
= = 2ra36"-Ta ly
(MO1) 2wf(a6)aW (21-8)
~e

1
where ta = /(7a) = /(a0); solving this for ta: o
\JL\A i
° ANGLE
#
T„ = s-ij—S ^ (M63)
-
. (21-9) OF TWIST PER INCH
**"a Fio. 21-4. Torque-tuist curve.

This equation serves for determining the stress-strain curve for shear
T = f(y) of a material if the moment-twist curve M = F(6) was observed
in The shearing stress t„ corresponding to the unit shear
a torsion test.
7„ = a6 being proportional to the quantity dependent on 6,

j, ^ (M62) = 6™ + ZM = <JP + 3IP (21-10)

Ta for various values of ya, i.e., a function t = f(y) may be computed by


drawing the tangents to the moment-twist curve (Fig. 21-4) in a few points,
by measuring the lengths of the two straight lines CP and XP and after
adding CP to 3 X AP. The shapes of the stress-strain curves for shear
T = f(y) have been studied by P. Ludwik1 and E. A. Davis2 for various
ductile metals in this manner and the observations in torsion tests
extended to large values of the unit shears 7 (which were computed for
large angles of twist 6 by defining 7 = tan <p, <p being the angle PQP'
(Fig. 21-1) under which an arc element QP' after distortion is inclined
with respect to the generatrix of the cylinder).
c. Length changes of permanently twisted round bars. C. W. MacGregor
on the suggestion of the author in 1933 made careful measurements (using
a Zeiss comparator) of the length of a round bar of soft copper after it had
1 Loc. cit.; also Ludwik, P., and R. Scheu, Stahl u. Eisen, vol. 45, p. 373, 1925.
2 J. Applied Phys., vol. 8, p. 213, 1937.
350 THEORY OF FLOW AND FRACTURE OF SOLIDS

been twisted near the fracture torque. He found a small permanent


increase of length in the twisted copper cylinder. E. A. Davis took
Vickers hardness readings along a radius in the cross section of a round
bar of copper after it had been severely twisted and cut in pieces. He
found that the hardness at the center of the circular cross section in the
axis of the bar corresponded to a hardness of copper which would have
been obtained by permanently straining copper in tension or compression
to a certain amount. Since the shearing stresses r according to the theory
described in a and b must vanish in the axis of the bar and the metal
around it in a cylindrical region of very small radius must be strained
only elastically, the copper should have shown a hardness when r = 0
of soft copper. These observations indicate that the center portion of a
cylinder of originally soft ductile metal, after it has been severely perma
nently twisted, is permanently strained in the axial direction. It must
be concluded that the distribution of shearing stresses r =
f(r6) consid
ered in a and b does not represent the complete system of stresses in round
bars after a considerable torsional overstrain1 but that a system of sec
ondary normal stresses in axial and possibly also in radial and tangential
direction must be present in addition to the distribution of the shearing
stresses t =
f(r6) deforming the metal.
J. H. Pointing2 was probably the first to call attention to a very small
lengthening of loaded wires when twisted in the elastic range. The
changes in length of permanently twisted metallic bars have been care
fully observed and reported by H. W. Swift* and R. L'Hermite4 and
G. Dawance.
Figures 21-5 and 21-6 reproduce some of the results of the torsion tests
made by Swift. Figure 21-5 shows the permanent axial extension of
twisted round bars in per cent for a 70-30 brass, aluminum, and copper
(<p

bar in function of the unit shear ya = tan <p angle by which a generatrix
of the cylinder tilted after twist).
is

That the torsion theory considered in a and and described in the first edition of
1

this book has a limited validity has not been pointed out formerly.
Proc. Roy. Soc. London, ser. A, vol. 82, pp. 546-559, 1909; ser. A, vol. 86, pp. 534-

561, 1912. He tested axially loaded steel piano and copper wires, twisting them in
the elastic range, and found that they showed a very small reversible increase of their
length when torque was applied. The lengthening was found to be proportional to
a

the square of the angle of twist.


Length Changes in Metals under Torsional Overstrain, Engineering, Apr.
3

4,

1947, pp. 253-257.


fitude theorique et experimentale sur la torsion composee, Comple rendu des
*

recherches effectuts in 1947, pp. 141-152 (Labor, du Batiment et des Travaux Publics,
Paris).
TORSION OF ROUND BAR 351

It is interesting to note from Fig. 21-6 that after a reversal of the torque the bar
shortens a little first and then again becomes longer. The effect of a preceding tor
sional overstrain in one direction, after the twisting moment is reversed, appears to
continue. According to Swift, there will be a tendency for accumulating slips in the
crystalline aggregate containing a large number of grains which contribute more to
an axial elongation than to the normal strains in other directions, and a fibrous struc
ture in rolled metals is in no way responsible for the lengthening effect which appears
primarily caused by the anisotropic behavior of a crystal aggregate under increasing
cold working.
Several facts may now be mentioned, at least, which were not considered in the
former theory (a and 6) which can explain the longitudinal extension of plastically
twisted bars.

1. The finite rotation of the principal directions of strain under a large simple shear
(see Sec. 13-4) may cause a simultaneous rotation of the principal directions of stress
because it is found that it will require less mechanical work to deform plastically an
element of a round bar if in addition to the system of shearing stresses t formerly con
sidered a normal stress a in the direction of the axis of the bar is also present through
which the shearing stress t doing the main work may be reduced while the octahedral
shearing stress at the plastic limit remains unchanged.
2. It is known that the density of severely cold worked metals decreases slightly
by an order of magnitude comparable with the elastic dilatation in volume produced
if a mean stress acts. If a round bar after considerable torsional overstrain should
carry a system of certain tensile and compression stresses in axial direction varying
with the radial distances from the axis but which would represent an equilibrium sys
tem of stress so that their resultant would still vanish, these stresses will produce in
the radial and tangential directions elastic strains of a magnitude varying with r caus
ing a secondary system of radial and tangential stresses. The small permanent
increase of the volume in the outer portions nearest to the surface of the bar which
were cold worked the most together with the variable elastic parts of the strains must
contribute to the increase of length of a severely twisted bar.
3. Under the increasing permanent strains the metal ceases to deform in the simple
manner postulated in all theories of isotropic flow. The permanent increase of length
of twisted bars is also due to the anisotropic way polycrystalline metals distort after
the strains increase to finite magnitudes. Although the theory of anisotropic plastic
352 THEORY OF FLOW AND FRACTURE OF SOLIDS

flow developed by R. Hill1 supports this viewpoint and could account solely for the
facts observed in the torsion tests by Swift and L'Hermite, probably the previously
mentioned causes participate in the observed effects.
1 A Theory of Yielding and Plastic Flow of
Anisotropic Metals, Proc. Roy. Soc.
London, ser. A, vol. 193, p. 281, 1948. Dorn, J. E., and A. J. Latter, Stress-strain
Relations for Finite Elastoplastic Deformations, J. Applied Mechanics, September,
1948.
CHAPTER 22

PLASTIC BENDING OF BARS


22-1. Bending of Bars with Arbitrary Law of Deformation. The theory
of bending of narrow bars, when the deformations do not obey Hooke's
law, may be stated in a relatively simple manner. We assume that the
bar has a cylindrical or prismatical shape with constant cross-sectional
area and that it is loaded by forces directed perpendicular to its longi
tudinal axis in one of the principal planes of inertia of the cross section.
The cross-sectional dimensions of the bar are assumed small, relative to
its length, so that the deformation due to shear may be neglected relative
to that due to normal stresses. Finally, cross sections having profiles,
composed at least in part of thin sections, and of unsymmetrical form
(angles or channels) will be excluded insofar as in such cases bending may
be combined with twisting.
The calculation utilizes certain methods of C. v. Bach and is based
on the assumptions made in the usual
; 2
theory of elastic bending that the cross
sections remain plane during bending. t \
The validity of these assumptions has p\ Z A

J ~b\
a\tSm~~
"""
been shown particularly by C. v. Bach1
i

on materials which do not obey Hooke's _


Jiq.
.. 22-1. Bending of bar.
law and possess no straight-line stress-
,
.

strain curve for tension or compression. These assumptions were also


confirmed by tests by Eugen Meyer* on wrought iron above the yield
point.
In order to determine the distribution of normal stresses over the
cross section (Fig. 22-1) at distance x from the end of the bar, the
a

stress-strain curve of the material for tension or compression must be


known from tests. We shall assume that this stress-strain curve has
been determined from a tension and a compression test in the shape of a
graphically (Fig. 22-2) or an analytically given function of the normal
stress a.
*=/(«) (22-1)
as dependent on the unit extension «.

Bach, C, and Baumann, R., " Elastizitat und Festigkeit," 9th ed., p. 259.
1

Z. Ver. deut. Ing., p. 197, 1908.


»

353
354 THEORY OF FLOW AND FRACTURE OF SOLIDS

It NN (Fig. 22-3a), the neutral axis,


is evident that there is an axis
along which the stresses and strains produced by bending vanish. If the
bar is bent, two neighboring cross sections x and x + dx are caused
to move so as to be slightly inclined to each other. A length dx (Fig.

Fig. 22-2. Stress-strain curve for Flo. 22-3. Cross-sectional area A and
bending. element A dx of a bar with distribution of
bending stresses a as depending on dis
tance v from neutral axis NN in section.

22-3c) at a distance i\ from the neutral axis undergoes thereby an exten


sion A dx; therefore the unit strain at a distance 77 is

A dx
1 —
dx
Since
A dx
dx
we have

(22-2)

In this is the radius of curvature of the elastic line of the bent bar.
p
The equilibrium of the forces acting on this element dx of the bar is
expressed by the two equations

j<rdA = 0 , (22-3)
fatj dA = M . (22-4)

In these equations dA represents an element of the cross section at a


distance and a the normal stress acting on this element.
y, The bending
moment which must act to hold these internal stresses in equilibrium is
denoted by M.
Instead of the element dA of the area A we may write also dA = /3(j») dri
if the width of the cross section at the point 77 is designated by b = f}(y).
PLASTIC BENDING OF BARS 355

Finally, if the variable jj is replaced according to Eq. (22-2) by the unit


strain t, Eq. (22-3) takes the form :

fm adA = p
[" /(e)/3(p«) de = 0 . (22-5)

Here a = i/i/p, «2 = v*/p are the absolute values of the unit strains in
the points of cross section farthest from the neutral axis. Since

iji + fit = h

the upper limit «2 in the integral is equal to a = - —


«i so that the limits
P
are determined by the values of «i and p. The functions and /3 may be /
taken as given and of the two quantities ei and p one can be chosen
arbitrarily.
Then the integral

f'jkMpt) dt = 0 (22-6)

becomes an equation for the other unknown quantity. By means of


this equation, function «i = ^(p), corresponding to a given shape of the
a
cross section and to a given law of deformation, is determined such that
even if no analytical expression exists it may be constructed point by
point by trial. (One needs only to choose a p for an assumed value of ei,
calculate the integral (22-5), and plot its values as obtained for different
values of p; the point where the integral is zero then gives the desired
value of p.)
h
If in this way «i = ^(p) and «2
= yi are determined, the maximum
P
stresses are given by <r\ = f(n), <ri = /(«2). For each pair of values ei,«2
(or a\,a) there corresponds a bending moment M which may be calculated
by (22-4), and curvature - = *2 •
Since the bending moment M
h "t"

a —
p

given function of the variable x, curvature 1/p may be found cor


is

a
a

responding to the value of M at the point x. Thus a certain curve — the


elastic line of the bar— defined by the differential equation
is

S-r
Since 1/p function of M
and hence of x, integration of this equation
is
a

gives the deflection of the elastic line as a function of x.


y

In the following sections few applications of the above derived rela


a

tions will be made.


350 THEORY OF FLOW AND FRACTURE OF SOLIDS

22-2. Pure Bending of a Bar with Rectangular Cross Section, a. To


Determine How the Bending Moment Varies with the Deflection. As in
this case M is independent of x, and b = 0(ij) is independent of 77, Eq.

(22-5) reduces to

If
f /(«) d* = 0

we introduce here the areas under the


. (22-8)

respective stress-strain curves for com


pression and for tension (Fig. 22-4),

- /(.)*,

jlm
.4, =

=
/„"/(«)*,

A
(22-9)

2
this condition then gives
Aj = A2. (22-10)
Fig. 22-4. Stress-strain a ~
One may then plot Ai and A2 as
curve

a
/(e) for bending.
function of the upper and lower limits,
respectively, of the integral (Fig. 22-5) and from this determine «j as

a
function of t\. The bending moment then equal to
is

M = bp2
j'j(*)*de. (22-11)

The angle which two cross sections at distance apart (Fig. 22-6) are
a

Fig. 22-5. Areas A\ and At as depending Fig. 22-6. Bar in pure bending.
on g.

inclined with respect to one another,


is

2c>
= - •
(22-12)
P

Since
vi = «ip vt = =
+

«2p vi (22 13)


h

v2
,
,

the curvature becomes

= and = '
~h~ —2h— (22-14)
*
p
PLASTIC BENDING OF BARS 357

The results of the calculation may be condensed into the two following
formulas:
Bending moment:

M
Jf"-ti
= Pa = /(«), de . (22-15)
Ui k2
~t~ ti)

Slope of the tangent:

* = •
(22-16)
^-±r^
<tt>

Since M and may easily be determined by a bending test, these formulas


permit convenient experimental test of the fundamentals of the calcula
a

tions. They give the moment M (or the load on the bar =

P
M/a) as
function of the slope of the elastic line the law of deformation of the

if
a

material known.
is

To Determine the Stress-Strain Curve from Bending Test. The above


b.

a
equations permit the solution of the inverse problem. It will be assumed
that from a bending test with a bar of rectangular cross section both
strains ti and a of the outermost fibers have been determined for series

a
of loads the same thing, for various values of the bending
or, what
P

is

moment M One may determine from these observations the


= Pa.
shape of the stress-strain curve of the material for tension and compres
sion as follows:
From Eq. (22-16) the slope of the tangent to the elastic line of the bent
bar given by
is

= («i + «) •
(22-17)
*

The position of the neutral given by (22-13) and (22-14)


is

axes
:

" = = ~
(22-18)
In the integral
f(e) df = 0, (22-19)
-ei

the limits ti and e: are to be looked on as functions of From these


it

<t>.

follows, therefore:

Lm dt = /(«,) + =
(22-20)*
0
I
I
h

In forming the derivatives of the definite integral, we proceed according to the


*

-
rule:

/(*,«) dx = dx +f(b,a) — (,,«)


^ ^
/
Ja
Ja

In Eq. (22-19), does not depend on the parameter </>,and for and we substitute
a

6
/

= —
u = ti
u

.
;
358 THEORY OF FLOW AND FRACTURE OF SOLIDS

or since,
ffl = -/( — *l) , 0-2
=
f{tt) ,

we get
at dt- = <ri At. (22-21)

The apparent meaning of this equation is that the increase of both


areas Av and Ait namely, en rf«i and <n da, must remain constantly equal,
as the load or moment increases.
The expression for bending moment [Eq. (22-15)] gives, if the value
<f>

of in (22-16) substituted, the following equation:


is

["
M2
=
M4>* -r /(«)«&• (22-22)

/
From this follows that
it
d_

d4>

or using Eqs. (22-21) and (22-16)

— = dt\ = -s-
+

ffidei

0

(72^*2 d<£
+
,

<Ti ] C2
I

, 2/i
2hdd>
= — =—
,
= -r —ffi —
x
,

,
• - ,
+
,

dei a«2 a«2 ""P


>

ITi + ffj >


l
(

we finally obtain

W -*£!_- IiL (W) (22-24)


ffl + C2 »<P
9

The right side of this equation known from observed test data, the
if,
is

<f>.

bending moment M known as a function of the slope From the


is

observed strains €i and e2, the ratio dti/da, and therefore using Eq.
(22-21), er2/cri may be determined for various values of the observed angle
<t>.

These expressions permit the calculation of the stresses <r\ and at


farthest from the neutral axis as functions of the strains or, in other words,
the construction of the stress-strain diagram for tension and compression.
'

Herbert, Mitt, u Forschungsarb. Ver. deut. Ing., vol. 89, 1910, used similar relations
1

in order to determine the stress-strain curve for cast iron in tension and compression
by means of bending tests on cast-iron bars. Cf. also Bach and Baumann, "Elasti-
zitat und Festigkeit," 1924; and J. Fritsche, Arbeitsgesetze bei elastisch-plastischer
Balkenbiegung, Z. angew. Math. Mechanik, No. 13, 1931; also Stahlbau, Nos. 16, 17.
1938; Nos. 14, 15, 1939; McCullouoh, G. H., An Experimental and Analytical
3,

Investigation of Creep in Bending, Trans. ASME, p. 55, 1933; Cozzone, F. P., Bend-
PLASTIC BENDING OF BARS 359

Example. Material with the same stress-strain curve for tension and
compression. For all degrees of deformation

«i = «2 , <r\
= f2 , (dei = dtj) . (22-25)

From Eq. (22-24) a\ may immediately be calculated:

'
(22-26)
I

These equations suggest the following rule for constructing the stress-
strain curve: in order to determine from a bending test the normal stress
<ricorresponding to a given unit strain ti
the curve representing the bending mo
M.
ment M as a function of the slope
<t>
M

is
f 9 ^^^^^"^
plotted (Fig. 22-7). The normal stress a\
equal to the projection PB of the tan
is

gent APof the moment curve M = F(<l>) M


upon the ordinate, increased by twice the
value of the bending moment and the sum
divided by bh?/2 = width, = height
0 c \ i
(b

of the rectangular cross section of a bar Fio. 22-7. Pure bending. Bend
bent by pure bending moment, = ing moment if
= F($); 2* angle
a

of inclination of elastic line of


I

length of the bar for which the angle gauge length


4>

t.

was observed).
22-3. Bar Subjected to Plastic Bending Having Idealized Stress-
Strain Curve Fig. 22-8. If bar of ductile metal loaded in bending
is
a

under increasing load, at a certain definite load in certain parts of the


bar the limit of plasticity will be reached. If the stress-strain curve for
the metal in tension and compression known in both the elastic and
is

plastic stages, the stresses inside a bent bar may easily be calculated
by the procedure as set forth in Sees. 22-1 and 22-2.
a. Initial Yield or Flow in Steel Bar. For a metal which has definite
a

yield point as, for example, soft annealed wrought iron, the stress-strain
curve may be replaced, for the above purpose of calculation, by three
straight lines. It thus assumed that the strains produced by plastic
is

ing Strength in the Plastic Range, J.Aeronaut. Sci., 1943, pp. 137-151 Jensen, V. P.,
;

Ultimate Strength of Reinforced Concrete Beams as Related to the Plasticity Ratio


of Concrete, Bull. Univ. Illinois, vol. 40, No. 44, 62 pp. 1943 (bending of beams of
concrete assuming a nonlinear distribution of stress); and Hrennikoff, A., Theory of
Inelastic Bending with Reference to Limit Design, Trans. ASCE, vol. 113, pp. 213-
268, 1948 (applications to statically indeterminate cases of bending beyond the yield
point).
300 THEORY OF FLOW AND FRACTURE OF SOLIDS

bending are not larger than a few per cent. If it is assumed that the yield
points for tension and compression are the same and that in the elastic
regions — «0 < t < «o, however, the strains may be calculated by taking
a constant value of the modulus of elasticity E = <r/t; these assumptions

if
correspond to a stress-strain curve
which satisfies the following conditions
(Fig. 22-8):
Wnsion
for e < —«o , <r = —<ro = const ,

for — = Et
/
COMPRESSION UNIT to < t < to , <r ,
ELONGATION
for «o < t , a = <r0
= const .

Fio. 22-8. Idealiied stresH-strain


In the inside of to a bar, subjected
curve for pure bending.
bending, which is stressed above the
yield stress, for certain parts of the bar two kinds of regions are to be
differentiated according to whether the strains are elastic or plastic (Fig.
22-9o). Likewise the elastic and plastic areas are to be differentiated in
different cross sections (Fig. 22-96). We consider a cross section of the
bar and take 17 equal to the distance from an arbitrary straight line, per
pendicular to the plane of bending (Fig. 22-9c). In a certain part of the

Fio. 22-0. Simply supported bar bent by single load beyond plastic limit.

cross section which we may designate by A' (Fig. 22-96), the bending
stress a is equal to the yield stress for tension <r0; a second part A" has
stresses equal to the yield stress in compression a = — o*o. In the middle
part A (Fig. 22-96), where the bending stresses have not reached the
yield point, the stress must satisfy the straight-line law:
617

a = a + (22-27)
.

We set S,S',S" equal to the static moments relative to the axis OO and
J,J',J" the moments of inertia of the above-mentioned areas A, A', A".
At the limits of the plastic zones = and = i\" the stress <r = +<r0,
v'

r\
v

or
+ br/' =
a

<r0
= -<ro
(22-28)
+ by"
a
PLASTIC BENDING OF BARS 361

From this follows that

a = —<ro
—+ 2<r0
(22-29)

v V V V

The condition of equilibrium


= 0
JV dA , (22-30)

when the above relations are introduced, takes the following form:

A -
a + S •
b = <ro(A" - A') . (22-31)

If, in this equation, a and b are replaced by the expression (22-29), there

results a function of and i\" . From this function the ordinates and

v'
rj'

ij", which separate the plastic areas of the bar from the elastic area, may
be calculated. If and are determined in this way the bending moment
a

may be calculated from

M = J<rij dA = Sa + Jb
+
(S' - S")<t0 (22-32)

An Example the Spread the Plastic Regions in a Bar Stressed Slightly above
of
of
b.

the Yield Point. As an example we


choose simply supported bar with tri
a

angular cross section of steel, loaded in


the middle by force P. The cross sec
a

tion of the bar an isosceles triangle


is

having and a height The


h.

base
a

length of the bar and the abscissa x


is
is
I

measured from the left end. The process


of calculation will be briefly outlined.
In the elastic region of deformation
the bending stresses are distributed in Flo. 22-1,. (a) Cross-sectional area and (6)
the cross sections according to the distribution of bending stress.
straight-line law:

' 12Px
^ ('-£) (22-33)

As the force P increases, the yield point first reached in the outermost compression
is

fiber = 0) of the middle cross section where x = 1/2, at a load:


(ij

P = P„
67
(22-34)

Since the stress at a load Po only a0/2 on the tension side = of the middle cross
(ij
is

h)

section, the bar begins to yield at first in the neighborhood of the point = on the
0
ri

compression side. At a load P > P„ there results at first only one plastic region in
the bar. This represented by the crosshatching in Fig. 22-lOa, its area being desig
is

nated by A".
In the part not crosshatched (the elastic part) of the cross section, the straight-
A

line law gives for the stress:


hr,

a = a + . (22-35)
302 THEORY OF FLOW AND FRACTURE OF SOLIDS

At the limits of A and A' we have for v


= ij"

—<r0 = a + bti" . (22-36)

Eq. (22-31) becomes when A' = 0,

Aa + Sb = <r0A" . (22-37)
Taking:

^ A S
2h 2h' Zh
(22~38)

and designating

o = aoa , b -^ . i/" -^ (22-39)

from the foregoing expressions we obtain:

" = ~ 1 " = '


(22"40)
2^-3^"'+ u"> 2 — 3»" +

This determines the stress distribution in the cross section ; for example, for

?<" < u < 1, a = (a + /3uVo ,

and for
0 < u < u" , a = —
CT0,

while the bending moment M becomes

* TT ~ "
«"• - 3u" + 2

(2IM1)

The above expressions hold as long as:

1
0 < u" < = 0.3661 . (22-42)
V^2~

As soon as u" becomes > 0.3661, P > 1.80 P0 and the bar begins to yield also on its
tension side. At a loading P > 1.80Po, there are two different regions inside the bent
bar in which the stresses are at the plastic limit.
In order to determine the stress distribution under this loading the following three
equations are available:
a + bi/' — at, \
a + = —<ro (22-43)
}
Aa+Sb = <r0(A" — A') . J
In this
f (h* —
v") ,„ _ ft,"' 5, r ,

S =
'(^- •
(22-44)

Using the above symbols as defined in (22-39) and taking

"'
U =
I ' =
X
' **
X ' (22_45)
PLASTIC BEXD1XG OF BARS 363

we obtain from the first two equations (22-43) :

*--%±X.>
u — u
f>--A-n-
u — u
' (22-46)

From the third we obtain:


u'» + u'u" + u"' = }i . (22-47)

This is the equation of an ellipse whose minor semiaxis a = 1 has the direction of the
bisector of the axes u' = u" and whose major semiaxis 6 = y/z. We consider only
that part of the ellipse where:

u' = l , „" = ^JT- - to u' - u" - -i= • (22-48)


•«
V2
To the one limiting point corresponds the point in the bar where the second plastic
region begins. If, however,

u' = u" ~ ' (22-49)


V2
there results a limiting case of bending of the bar. In this case both the above differ
ent plastic regions touch each other in the middle cross section of the bar. This
occurs when P = 2 34/Y
In the middle portion of the bar, where it is strained plastically on the tension as well
as on the compression side, as long as I.8OP0 < P < 2.34Po the bending moment is

On the basis of the preceding formula the limits of the plastic areas for P = I.8QP0
and for P = 2.34/>0 are represented to scale in Figs. 22-11 and 22-12. In these the
plastic parts of the bar are represented by the crosshatching.

PLASTIC i.P-/.e0Po
22-11.

22-12.

Figs. 22-11 and 22-12. Limits of plastic regions in a freely supported bar of triangular
cross section bent by single load P at center. The boundaries of the plastic regions are
indicated in true scale for two loads. Fig. 22-11 (above): Load P = I.8OP0 just when
lower side of bar starts to yield. Fig. 22-12 (below): P = 2.34P0 when whole middle sec
tion yields. (Po is the load under which the bar just starts to yield at upper side.)

It is clear that the above calculation insofar as it relates to the spread of the plastic
regions under the larger loads will decrease in accuracy with the increasing load
(P = 2.34P0). At the higher loads the validity of the simple assumptions made
relative to the shape of the stress-strain curve a = f(e), (viz., that the yield stresses
are independent of the strains t) on which the calculation is based, do not hold accur
ately. How these assumptions may be improved is indicated in the following section.

22-4. Plastic Bending Considering Work Hardening According to


Idealized Stress-Strain Curve of Fig. 22-13. The assumptions regarding
364 THEORY OF FLOW AND FRACTURE OF SOLIDS

the shape of a stress-strain curve, on which the treatment of Sec. 22-3 is


based, do not take into account the increase in yield stresses or the strength
ening of the material (work hardening) which occurs for most ductile
however, no difficulty in taking

is,
metals with increasing strain. There
this factor into account. How this can be
done will be illustrated by the following ex
ample of the plastic bending of an iron bar
with a rectangular cross section.

a. Stress-Strain Curve, According to Fig. 22-13.


The calculation can be based on a stress-strain curve
composed of three straight lines as shown in Fig.
22-13:

for —eo >

«
Fig.

,
Idealized stress-
22-13.
for —to < < a = a'l
«o

«
strain curve for steel, taking

,
//
for to < +

<
into account work hardening.

,
If we introduce here instead of the constant <r0othe yield stress for tension c0 and con
sider that at the break of the curve — to the stress

is
e

=
+
x'«o Too ir1 «0 (22-51)
,
we then have

-(>-£) (22-52)

The stress-strain curve thus determined by the following three constants depending
is

on the material: the yield stress <t0,


which the same for tension and com
is

pression, a constant <r' for the elastic,


and a constant a" for the plastic strain
equal to the modulus of elasticity
is

(a'
of the material E, a" a measure of the
is

increase in yield stresses under increas


ing strain, i.e., of the rate-of-strain
hardening).
The length of the bar designated
-b-
is

by the width of the rectangular cross


Fia. 22-14. Distribution of bending stresses
I,

section by the height by and the


h,

in cross section.
6,

load of the bar by P.


On account of the symmetrical distribution of bending stresses, not necessary in
is
it

this case to determine the position of the neutral axis. The stresses in the cross sec
tion, distributed according to Fig. 22-14, give the following bending moment:

M - 2(P„, P„t + P,v,)


+

(22-53)
.

In this Pi, Pi, P, represent the resultant values of force produced by those portions of
PLASTIC BENDING OF BARS 365

the stress diagram (Fig. 22-14) which are bounded by the straight lines shown; and iji,
vi, Vi represent their moment arms relative to neutral axis (Fig. 22-14) :l

P* = *J> - -

,o)
v, :2

+
do)
(22-54)

j
_

y
(h +

(h
<r„6<r" „„)

3
(The symbols relate to Figs. 22-13 and 22-14.) If we put

u = —

^
(22-55)

the bending moment becomes

M _
Px- ^ - PJv

if -
(22-56)


The bar begins to yield under a load:

P.-^-°- (22-57)

The above equation for M holds for The variables u and are to be con
P

> P0.

p
sidered as functions of x. In the middle of the bar we have for x = 1/2, = P/P0.

v
If this value of designated by v0 = P/Po, the equation of the boundary curve of the
is
v

plastic region determined from the following relation:


is

r,—,.+ga—«)'o «n.
+
fa

" « (22.58)
2i»0 4»0
J
L

At a certain load P, where P > P0, at those points of the cross section, where u =

1
and = the yield point in the fibers farthest from the neutral axis just reached.
is
1,
»

The abscissas of these cross sections may be obtained from (22-58) by substituting
=
t>

it - and
1

1
:

x - *, - ± (22-59)
.

Pi determined as follows: let equal the stress along the edge and n the corre
1

is

<ri

sponding strain for — Then P% = _ _ Since


^
ij

^.

«o _ 2«i
1

i?o
A
P

we have «i = A«0/2i)0 " (<m


— <roo)A", from which we obtain

" = <roO""
; \

-v w.
/
A

-" — ~ •
1

Introducing this in the above expression for P we obtain the value given in Eq.
2,

(22-54).
366 THEORY OF FLOW AND FRACTURE OF SOLIDS

In the plastically bent part of the bar xt < x < 1/2, the maximum stress at the edge
of the cross section having the abscissa x is given by

= ±k„ +</'(«, - *„)] = (22-00)


+<r„[l +£-'^Jf-)]
The shape of the elastic line of the bent bar for that portion where the strains are
purely clastic is determined from the well-known equation:

d*y M Px
y= Pxx
C,X+C" (»<*<*)■
dx~>= ±JE= -*JE' -12JE +

In the plastically bent part of the bar it is determined by the formula:

1 < «o
P V lo

If the strain e0 in this formula be replaced by a0/E and the ordinate t/0 of the boundary
curve of the plastic region by >jo = hu/2 and if the curvature 1/p is replaced by
—d*y/dx', we obtain for the plastically bent part (xi < x < 1/2) the following differ
ential equation of the elastic line:

v
dx1 Ehu

As will be recognized from the formulas and the expressions for the limiting curves
of the plastic regions, the elastic region in the middle part of the bar, at sufficiently
large loading P, shrinks up into a narrow strip.
In a bar which is plastically bent under heavy loading, it is possible to differentiate
three different parts: an elastic part, having pure elastic tensions (0 < x < Xi), an
elastic-plastic part (xi < x < xs), and a completely plastic part (xj < x < 1/2).
Under high loadings, producing severe yielding, in the transition part of the bar
Xi < x < Xi, u and v may usually be replaced by the simpler expressions:

and in the plastic part these may be replaced by

3 , o" <t"u 2Px 3 m „,.

The deflection of a bent bar at high loads may be calculated with sufficient accuracy
in that portion where 0 < x < x2 from the differential equation:

y"'-m- ^4)
In the plastic part x2 < x < 1/2, it may be calculated from the equation:

The limiting curves of the plastic regions in a severely bent bar may be
observed experimentally in various ways. In bars of very soft iron the
PLASTIC BENDING OF BARS 367

limits of the plastic regions may be indicated by suitable etching com


bined with previous thermal treatment of the test bar. For this purpose
the etching method of A. Fry already referred to above or etching of
the iron by heating to its recrystallization temperature is of value.1

Fio. 22-15. Etching of a section through an old boiler plate showing flow lines and cracks.
(According to B. Strauss and A. Fry.)

A few photographs of the markings obtained using Fry's method in the


longitudinal sections and cross sections of plastically bent steel sheets
are shown in Figs. 22-15 to 22-18. Figures 22-15 and 22-16 show two
longitudinal sections in the plane of bending of a steel plate. Figure
22-17 shows a section perpendicular to the plane of bending. The posi-
Fio. 22-16.

Fio. 22-17.
Figs. 22-16 and 22-17. Liiders' lines in sections perpendicular and parallel to axis of a
steam boiler. The flow lines were produced by the bending of the boiler plate in the
bending machine. (According to B. Strauss and A. Fry.)

tion of the flow layers is indicated in the sketch of Fig. 22-18. As will
be seen, the flow layers in a bent steel sheet practically coincide with
the surfaces of maximum shear. In longitudinal cross sections of a
wrought-iron bar bent beyond the plastic limit, the shape of boundary
lines of the plastic region could be demonstrated using Fry's etching
1 Fischer, Fk. P., Krupp. Monatsh., vol. 4, p. 77, 1923.
3f>8 THEORY OF FLOW AND FRACTURE OF SOLIDS

method. The boundaries of the dark surfaces in the etchings were prac
tically identical with the shape calculated by the above methods for the
case of a simply supported bar loaded in the middle.
Using the second method, the steel bar
stressed above the yield point is heated
for some time at 650 to 750°C. The
annealing treatment at this temperature
induces in the stressed steel a rearrange
ment of its microstructure, and, accord
Fio. 22-18. Flow lines in a plate
produced by bending. ing to a well-known law, under recrystal-
lization the largest crystal grains are
formed at the limits of the plastic region, which may then be made visible
by the usual etching process (e.g., by means of copper-ammonium-chloride).
6. Other Examples of Plastically Deformed Regions in Bent Bars. In calculating the
shape of the boundary curves of the plastic region it was assumed that the stress-strain
curve of the material is represented by the broken line of Fig. 22-8.
1. Simply Supported Bar with a Single Load in the Middle. If the material
begins to flow under constant stress, the constant a" — 0 in formula (22-55). From
formulas (22-56) and (22-55), the bending moment becomes

M =
Px Pol(Z - w)
(22-66)
8

from which the equation of the boundary of the plastic area follows:

U* = 3 - 4Px
PJ.
'
(22-67)

The plastic area is therefore bounded by two parabolas, which shift along the bar as
the load increases. The limiting position which these parabolas can take up is that
position in which their vertices just touch in the middle of the bar. The force is then
/mas ~ uTd/*'
2. Simply- Supported Bar Loaded with a Uniformly Distributed Load. The
length of the bar is /; the rectangular cross section has a width 6 and a height h, while
the load per unit length of the bar is p. The bar begins to yield under a load p0:

ibh*«<,
Po (22-68)
31'

If p > po, there occur in the middle of the bar two plastic regions which are bounded
by two branches of a hyperbola. The bound
ary line of the plastic zone has the equation: 0.29891 p

A1 -zr.
= 1- (22-69) mummm
In this the semiaxes are ± <^^g»p- w
Po 2p Fio. 22-19. Plastic regions in a bar
bent by uniformly distributed load p.
(22-70)

at a load p = 3p0/2, the hyperbolas coincide with their asymptotes (Fig. 22-19)
PLASTIC BENDING OF BARS 369

3. Bar Having Built-in Ends with Uniformly Distributed Load. In this


case the plastic regions are bounded by elliptical arcs.
22-6. Load-carrying Capacity of Steel Bars Bent in Plastic Range of Strains.
During the last two decades practical significance has been attributed by civil engi
neers to the term "load-carrying capacity" in the plastic range of strains of steel
structures. Suppose that we consider the ratio dP/dy of the increase of a significant
load P deflecting a steel girder or beam to the increment of its largest static deflec
tion y. In structural members of constructions low-carbon steels are used character
ized in their tensile stress-strain curve by a comparatively large yield point elongation
(see Chap. 19) or a range of the permanent strains of several per cent in which the
load will practically stay at a constant value and beyond which the load will slowly
increase until it reaches its greatest value in a comparatively flat maximum. For
such very ductile steels the value of the ratio a" /a' of the constant a" introduced
in Eq. (22-51) for measuring the slope of the true tensile stress-strain curve within
the range of the yield point elongation and a' = E measuring the slope in the elas
tic range is about one-thousandth (a" /a' = ~<r0/£ = 0.001) and may be assumed
equal to zero. In other words within this range of strains the stress-strain curve
may be represented by the idealized diagram of Fig. 22-8, Sec. 22-3. It may be
seen that under these circumstances the ratio dP/dy, for example, for a bar of
mild steel simply supported at both ends and loaded by a concentrated force P
in the middle of its span / will decrease rapidly with the deflection y after the
yield point has been reached in the center portions of the bar approaching the
values dP/dy = 0.
One verifies easily that a beam of a material yielding according to the idealized
stress-strain curve of Fig. 22-8 and having a rectan
gular cross section (6 width, h height) and a length I
between supports, which would start to yield in the
middle section under a bending moment Mo and a
load P0

T " T"
Pol vobh*
M> = (22-71)

where <r0 is the yield stress in tension, would be


deflected subsequently in its permanently over
strained middle portion by bending moments
Fio. 22-20. Bending moment
M "
&) M

TT 0 (22-72) as a function of strain «i.

(this last equation being valid only for «i > «o, see Fig. 22-20) where «0 = <ro/E is the
elastic strain corresponding to the yield stress <r0and «i denotes the total strain (sum
of elastic and of permanent strain) in the extreme fibers of the rectangular cross sec
tions. If ij denotes again the distance of a point in the cross section from the neutral
axis and ijo the distance of a fiber just starting to yield,

e0i7 . h
(22-73)
1)0 &

the bending moment M is also expressed by

M "
TV ~
3^J (22-74)
370 THEORY OF FLOW AND FRACTURE OF SOLIDS

When >)oin the most stressed middle section of the beam approaches the value zero
the strain «i will increase indefinitely and the maximum bending moment in the bar
will reach its largest possible value

At this load
M „» =
^ =


(22-75)

~* (22-76)

the bar deflects indefinitely and its equilibrium will cease to be stable; also

,— -o.
That the deflections y of the bar will rapidly increase follows also from the differential
equation of the elastic line derived from Eq. (22-61) assuming <r" = 0 within the
permanently overstrained portion of the beam (where rj0 < h/2) and which takes now
the simple form

showing that the curvature 1/p = » when tj0 = 0.


Thus we see that Pu, = 3P0/2 or a load which is 50 per cent larger than the load
Pa under which the beam just started to yield defines the load-carrying capacity of a
beam of rectangular cross section of a very soft steel yielding according to the idealized
stress-strain curve of Fig. 22-8, Sec. 22-3.
After the stress in the stress-strain curve of a mild steel starts to increase beyond the
yield point elongation, this will tend again to stabilize the equilibrium in bending; but
even if sufficient allowance is given to strain hardening in an actual stress-strain dia
gram conditions of this kind must be considered as dangerous, leading to severe local
overstraining in the material in the middle portion of the bar.1
1 The danger of collapse of important members in structures supporting buildings
subjected to bending in case of exposure to heat due to a conflagration, or to the
action of excessive winds, etc., may be judged by estimating the load-carrying capacity
of important elements beyond the elastic range of strains. The reader is referred in
this connection to a valuable summary of considerations for judging the safety of
steel beams overstrained beyond the yield point in bridges and structures in a paper
by F. Bleich, La Ductilite de l'acier (L'Ossaiure mttallique, No. 2, pp. 93-105, 1934),
in which statically indeterminate cases of bending in continuous beams are also treated.
This author points out that the theoretical load-carrying capacity Pn*x, for example,
of an I beam of standard profile is only 16 per cent greater than Pa, the load under
which the flanges yield first, the web contributing so little to the areas in the section
carrying the major bending stresses in the beam. The estimation of the load-carrying
capacity of a beam bent in the plastic range of strains described in the text above
furnishes the limiting loads well on the safe side. It need scarcely be added that it
represents an oversimplified way of treating the case of plastic bending in bars when
the permanent strains become of finite magnitude. It does not take account of (1)
the lateral distortion of the cross sections of the bar, (2) the fact that dPy/dx* is not
equal to 1/p when the deflections y increase to larger values, and (3) the strain har
dening of the steel, etc.
CHAPTER 23

BUCKLING AFTER THE YIELD POINT IS EXCEEDED


It was shown by von Karman1 that the theory of bending of bars may
be extended to those cases in which bars axially loaded in compression
are stressed beyond the yield point. If a bar is loaded in compression its
equilibrium may, as is well known, become unstable if the slenderness
ratio is sufficiently large. The bar then tends to buckle. Now, if the
stresses which occur when the bar is straight and subjected to axial
compression reach the yield point, it is clear that during bending the
compressive stresses on the convex side of the bar are decreased (by the

ai \ <r=f(fj

Fio. 23-1. Idealized stress-strain curve for Flo. 23-2. Bending stresses a in bar start-
buckling after the yield point is exceeded. ing to buckle.

tensile stresses set up by bending), while those on the concave side are
increased. In this consideration the treatment will be based on a stress-
strain curve (Fig. 23-1), such as was assumed in Sec. 22-4. The bar
has a rectangular cross section of width b and height h, a length I, and is
loaded axially by a compressive force P to an average stress o-*, which is
higher than the yield stress <r0 of the material. It is desired to find the
critical value of compressive stress <rc or the limiting force Pc = bhae, at
which the equilibrium of the bar in its straight form becomes unstable,
and buckling results.
In Fig. 23-2 the stress distribution in a bar is represented. Before
bending, the stress distribution would follow the horizontal straight line
a = <re. After bending, the stress distribution is represented by the
1 Untersuchungen fiber Knickfcstigkeit, Mitt. u. Forschungsarb. Ver. deul. Ing.,
Heft 81, Berlin, 1909.
371
372 THEORY OF FLOW AND FRACTURE OF SOLIDS

broken line AiOA2. The equilibrium of the stresses requires that


-
17)i <r2r/2 . (23-1)

Since the cross section remains plane during bending, using the symbols
of Eq. (22-2), page 354, and of Eq. (22-51), page 364, we have

«! =
— = —/ «2
= — = — <ri
= > a2 = •
(23-2 )
" P * P p p

Moreover,
Vi + V*
= h. (23-3)

From these expressions it follows that

<r JJi* — <T »Jj*

and

— h V7~> hV7

The bending moment is equal to

M =
| (^7>!* + <rji?,*) , (23-5)

or applying Eq. (23-4) we have

M - bh3 <rV
W)
(V? + v70'-
Using the symbols

(J being the smallest moment of inertia of the cross section about a


gravity axis) we thus obtain the differential equation for the deflection
of the bar:

1= dy* — M _ Py
(23-8)
p dx2 JE0 JE0'
From this it follows that:
I p rp~~
y = A sin x + B cos -^j^r a; .
(23-9)
yj

Using the boundary conditions:

*=±5>
V-0, P =
nVJ|', (n=l,2, ...)• (23-10)
PLASTIC BUCKLING 373

The smallest load at which the bar buckles is therefore

Pc=**J-p- (23-11)

This formula is of the same type as Euler's formula for elastic buckling,
namely,
,JE
Pc -T-
It is seen that the value of modulus of elasticity E given by Euler's
formula has simply to be replaced by the value E0, given by Eq. (23-7).
The formula (23-7) derived for rectangular cross section may be
immediately generalized for any arbitrary cross section. From the con
dition of equilibrium fa dA = 0, using

<r = — r\ , or a = — ij ,
Vi Vt

the following equation results:

^1
/ vdA =^ / vdA.
Vi Jo V* Jo
Putting:

<7l
= J (7j = J
P P

the former equation simplifies to

by means of which condition an axis in the cross section is determined


along which the stresses do not change during the first instant of buck
ling. Putting Si, St equal to the static moments and J\, equal to the Ji
moments of inertia of the two halves of the cross section relative to this
axis, and = J J\
+ ■/» equal to the moment of inertia of the complete
cross section, the bending moment becomes

.1/
-/
- / andr,
= *-dl +
Vi
°jll
Vt
= I (a>Jl
P
+ a"J,) = l-JE0.
P

From this the value E0, from which the buckling load P* may be calcu
lated, becomes

E0 = a'J;+a"J*. (23-12)

A peculiarity of the buckling process, accompanying the exceeding of


the yield point, which von Karman has especially emphasized, is the
374 THEORY OF FLOW AND FRACTURE OF SOLIDS

decrease in the axial compressive force during bending. If one plots the
maximum deflection of a bar during buckling as abscissas and the com
pressive load P at which the bar bows out, as ordinates, for various values
of the eccentricity of the axial load and for elastic buckling the series of

DEFLECTION '— *» DEFLECTION-*-


Fig. 23-3. Etastic buckling. Load deflec- Fig. 23-4. Plastic buckling. Load deflec
tion curves for various eccentricities of load. tion curves for various eccentricities of load.
{According to Kdrmdn.)

curves shown in Fig. 23-3 is obtained. All curves have for their common
asymptote the horizontal straight line which corresponds to Euler's
critical load, Pc = rUE/l*.
With buckling, accompanied by average stresses exceeding the yield
point, one obtains a series of similar curves shown in Fig. 23-4 with the

Fig. 23-5. Buckling loads for various ratios of slenderness l/i.

difference that now with increasing eccentricities the axial maximum


compressive loads which produce buckling decrease considerably.
If one plots the compressive stress at the instant of buckling against
the so-called "slenderness ratio" l/i (where i = v J/
A equals the radius
of gyration of the cross section), there is obtained according to Karm&n's
tests a line such as is represented by the heavy line in Fig. 23-5. The
PLASTIC BUCKLING 375

portion of the curve AB corresponds to elastic buckling as given by


Euler's buckling formula:

f. = * p

The portion BC corresponds to buckling after the yield point is exceeded.1


Considerable work on the buckling of straight bars and columns on
small models of frames of steel structures beyond the yield point has
recently been carried out at the Engineering Laboratory of the University
of Cambridge, England, by Professor J. F. Baker and his associates under
the sponsorship of the British Welding Research Association, London, to
which reference is made.2
1 Relative to the practical application of the results of the theory of buckling in
bridge design and in proportioning structural compression members cf. W. Gehler,
"Vorschlag einer Gebrauchsformel fur Knickung," Mitt. d. Normenaussc.husses d.
deutschen Industrie, vol. 2, Nos. 11, 12, 15, November, 1923; in Bauingenieur, 1923.
Relative to buckling of a bar of unelastic material such as concrete and reinforced
concrete, cf. Bach and Baumann, "Elastizitat und Festigkeit," 9th ed., Verlag
Julius Springer, Berlin, 1924; and E. Morsch, "Der Eisenbetonbau, seine Anwendung
und Theorie," 5th ed., Stuttgart. Lately F. W. Geckeler in Z. angew. Math.
Mechanik., 1928, has calculated the buckling load for the case of inelastic buckling of
a thin-walled cylinder, which is loaded in an axial direction. The buckling of such a
cylinder is accompanied by symmetrical corrugations or waves.
An extension of the work of v. Karman by H. M. Westeroaard and W. R.
Osgood, Strength of Steel-columns, Trans. ASME, p. 65, 1928, should be mentioned
here. These authors give the theory of steel columns which are stressed beyond the
proportional limit and which are eccentrically loaded or initially curved.
2 Baker, J. F., and
J. W. Roderick, Investigation into the Behavior of Welded
Rigid Frame Structures, Trans. Brit. Inst. Welding, vol. 1, No. 4, 1938; vol. 3, No. 2,
1940; vol. 5, No. 3, 1942; 4th report, British Welding Research Association, 1945;
"Welding Research," vol. 2, No. 1, February; No. 4, August, 1948; and Roderick,
J. W., Phil. Mag., ser. 7, vol. 39, p. 529, 1948.
Concerning the theory of buckling of frameworks, cf. also Kavanagh, J. C.,
"Instability of Plane Truss Frameworks," Dissertation, New York University, Apr. 1,
1948, printed by Column Research Council, Engineering Foundation, New York, con
taining an extended bibliography on this subject.
Part III
THE ELASTIC, THE VERY VISCOUS,
THE IDEALLY PLASTIC SUBSTANCE, AND
SOME OF THEIR GENERALIZATIONS
SPECIAL PROBLEMS OF THE IDEALLY
PLASTIC SUBSTANCE
CHAPTER 24

SYNTHESIS OF SMALL ELASTIC AND PERMANENT STRAINS


As mentioned in the introductory Chaps. 1 and 2, in the study of the
permanent deformations of the ductile metals, the natural rocks, or cer
tain artificial materials, the simplest possible forms of the assumptions
concerning the mechanical properties of materials have frequently proved
to be of practical value. With the aim in mind of developing means for
predicting the distributions of stresses or strains in deforming bodies, the
theories of a number of special solids will be discussed in the subsequent
chapters by ascribing to them certain well-defined ideal properties.
The strains in the elements of a body yielding under a system of forces
in general have a recoverable, an elastic, and a permanent portion. In
many applications the elastic strains can be neglected entirely and only
the permanent strains need be considered. In other cases, however, this
is not permissible.
In this chapter a theory of small distortions is worked out for a material
which has a sharply defined limit of plasticity, assuming that the strains
in the body are purely elastic as long as the stresses have not reached the
limit of plasticity and that they may increase permanently when this
limit is reached. Under certain loads there will be portions strained
elastically and other portions distorted permanently in the stressed body.
In the latter regions the strains will also have elastic portions. A soft
steel under moderately large strains may be considered as an example of a
material which distorts in this manner.
Experiments with mild steels and with other metals, furthermore, sup
port an assumption, which will be made, namely, that the total strains
in the permanently deformed zones are the sums of the elastic and the
plastic strains. The resultant strain tensor at a point of a body is the
sum of an elastic and of a plastic strain tensor. The components of
the elastic strain tensor change proportionally with the components of the
stress tensor if Hooke's generalized law of elasticity is assumed. It is
understood that the elastic constants or moduli of the substance remain
unchanged during the flow — a fact which is verified by experiments in
many materials.
With regard to the orientation of the principal axes of the resultant
strain tensor relative to those of the stress tensor or relative to the direc
379
380 THEORY OF FLOW AND FRACTURE OF SOLIDS

tions of the principal stresses we have to distinguish two important cases


depending on whether the principal axes of the resultant strain tensor con
sisting of the sum of the elastic and the plastic strains coincide or do not
coincide with the principal stress directions. In the first case it will
frequently be sufficient to introduce elastic and plastic stress-strain rela
tions in finite form, while in the second case these will have to be assumed
for infinitesimal increments of the strains. It is important, however, to
add that it will be necessary to
consider in certain cases infinites
imal increments of strain in practi
cal examples when the components
of strain are very small in magni
tude. Stress-strain relations for
the increments of strain must be
used in general also as a conse
quence of large strains. Cases in
which the permanent strains become
finite will not be treated in this
chapter, however.
We shall consider the first case
and shall assume that the result
ant strain tensor is coaxial with
the stress tensor and remains so
during the
deformation.1 The
second case will be treated later in

Flo. 24-1. Simeon Denis Poisson (1781- Chap. 29.


1840). (Courtesy of Pierre Gauja, Instilut In order to express the stress-
de France, Paris.)
strain relations we shall denote the
elastic and the permanent portions of the strains by one and by two primes.
The components of the total or resultant strain will have no dashes. t»
denotes a unit strain, ynm a unit shear, an a normal, and t„„ a shearing
stress, E the modulus of elasticity, v' and v" Poisson's ratio for the elastic
and the permanent strains, respectively, G = E/2{\ + v') the modulus
of rigidity, and K = E/Z(\ — 2v') the bulk modulus of the material.
In accordance with one of the assumptions made, we shall assume that
the components of the resultant strains are the sums of elastic and of
permanent components:

U = u' + ••• 7v.


= 7y.' + 7v.", • • •
• (24-1)

1 The developments in this chapter are partly based on papers by H. Hencky and

A. A. Ilyushin.
SYNTHESIS OF STRAINS 381

For the components of the elastic strains we assume that Hooke's law
is valid:

t, =
E G

«„
= ' 7., /
7« =
75-
>
f
~g G
, at — v'(a, + <Xy) f Till
'
e<
E G

If «' denotes the mean elastic strain and a the mean stress

, t, + ty + «x' v, + <ty + °, <n. n\


t = I <T =
g g

the sum of the first three Eqs. (24-2) indicates that the mean strain t' is
proportional to the mean stress

t, = = <r— = fc
(J_-^> (2^4)

where the constant A; = 1/3/C.


A second form for the elastic stress-strain relations (24-2) is of impor
tance. After subtracting the mean strain t' from the unit strains «x',«„',e,'
in the Eqs. (24-2) and after using Eq. (24-4) they may be expressed also
as follows:
/ / ,x f' *Yy,
~
Ty%
*' *
2G 2 2G

• " ' (24"J)



~2~ ~2G''
*
/ _ / _ "t —
g ' 7,y' — T,„
2(? 2 2G

The six components of strain appearing on the left sides of Eqs. (24-5)
and similarly the six components of stress on the right sides constitute
the components of a deviator of strain and of stress, which symbolically
may be written as follows (see Chap. 14) :

\tl e
2 2

D > = \ t' — 1^
t-'
"
\

y» 7»' u, _ (24-6)
ti

)
)
2

— a
\<TZ Try T«
Ti/, <ry
— * Tyt
T,, a, —
Txy a)
382 THEORY OF FLOW AND FRACTURE OF SOLIDS

Since according to Eqs. (24-5) the six components of D, multiplied by the


factor 1/2G are equal to the six components of D,, these two deviators
must not only be coaxial tensors but their Mohr circles must also be
two geometrically similar groups of circles. Thus Hooke's law is also
expressed in condensed form by the two equations

D'' = ' *' = ka (24"7)

in which the first equation refers to the deviatoral part of the strains
symbolically expressed by D,- or to a distortion without a change of
volume and the second equation to the change in volume.
A third form of the three equations for the unit strains is also

-to + ^^-'JL+JL^,
• • •
etc (24-8)

We now have to consider the permanent portions of the strains.


Strictly speaking, we can establish relations connecting the components
of stress and the components of the instantaneous rates with which the
permanent portions of strain change with the time t only in a material
which deforms permanently. According to what was said in Chap. 16,
the rates of permanent strain must be expressed by six equations of the
form

7„, = + v tv, • - - 1
2(1 )<t> ,

where the dot above t,", . . . yv,", . . . denotes the derivative with
respect to the time t, or we may express these relations for the infinitesimal
increments of the six components of permanent strain during the time
clement dt as follows:

du" = d+'l: - *"(*, + a.)] ,


• • •
\ K**~M)
dyv," = 2(1 + v") d+"r» ,
- - .
j

satisfying the rules of flow stated in Chap. 16.


We may think, for example, of a distortion during which the permanent
strains increase only by very small amounts which are comparable with
the elastic strains while the components of stress in the ideally plastic
substance remain at their constant values at the yield limit in every point
of the material. The principal directions of stress will then not rotate
appreciably relative to the strained elements of the substance. Assum
ing under these circumstances steady slow flow, we may replace the incre
ments dt", . . . , dyvt", . . . , d<tt>" by the small values tx", . . . ,
SYNTHESIS OF STRAINS 383

7v«", ■ ■ . , <t>" themselves and can introduce instead of Eqs. (24-9a)


plastic stress-strain relations of the form

«," - y"(cv + a,)]


H'\c,
*"k, - *"(». + »,)]
= , yy," = 2(1 + V")<t>"rv
y„"
*"[a, -
= = 2(1
H" , + v")4>"r„ , \ (24-10)
u" = + »,)] yz," = 2(1 + *")*"»>
r"(<r. , 'n/ ,

where <£" is a variable quantity.


In order to express the equations for
the resultant strains in symmetric

<f>'
\/E
form, let us write for in Eqs. (24-2). =
After adding Eqs. (24-2)
and (24-10) we obtain, for example, for tz and for yt.:

(24-11)
7.. = 7,.' + y„." = 2(0' + «")
[l + "'^y] r„ .

With the abbreviations

"if-
= =
%
+ *">
*'

<24-lla>
y
*

X
we can rewrite the stress-strain relations for the components of the
resultant strain tensor as follows:

«i = — v{ay y„, = 2(1


4>W* + <r,)] + v)4>tv1
,

ty = <t>[ay
— v(a, + <r,)] yzz = 2(1 + v)<$n,t (24-12)
,

t, = 4>[<r,

v(<rz + <r„)] y^y = 2(1 + v)^^
.
,

<f>

,v' ,v" are material constants. and are variable functions of the
$

<t>"

character of reciprocal moduli. The quantity interpreted as variable


is

a
v

Poisson's ratio.
In the ductile metals the permanent strains usually contribute little
to a change in volume. Assuming an incompressible material Poisson's
ratio for the permanent portions of the strain may be taken equal to
„" = y2. The factor 2(1 + v")<t>" appearing in Eqs. (24-10) for the unit
shears later to be denoted by if/" = 2(1 v")4>" then becomes equal to
+

= For most ductile metals may be assumed equal to 0.3.


v'

if/" 3<f>".

Assuming in the following these particular values = 0.3, v" =


Yi

the
v'

permanent strains do not contribute to the dilatation of volume, and for


small strains:
«," + «„" + u" = . (24-13)
0

We may add that, under the preceding assumptions, the elastic, plastic,
and the resultant strain tensors are coaxial with the stress tensor and
remain so during a small permanent distortion.
384 THEORY OF FLOW AND FRACTURE OF SOLIDS

It may be worth mentioning that the stress-strain relations in their


form defined in Eqs. (24-10) may apply to more general cases of flow than
were considered above. They may, for example, be used for expressing
stress-strain relations in a strain hardening metal, in which the flow
stresses increase with the permanent strains. They are valid under the
following restrictive assumptions: (1) the permanent strains are com
paratively small, (2) the principal directions of stress do not rotate
appreciably relative to the elements of metal deformed, and (3) the two
ratios of the principal stresses <ri/a\ and <n/<r\ remain at constant
values while the principal stresses <t\,<ri,<t3 and the components <r,,<ry, . . .
increase with the strains t,",ty", ....
How under these restrictive
assumptions for a strain hardening ductile metal the function 4>" may be
expressible will be shown in Sec. 28.-3. 1

Use of Eqs. (24-10) and (24-12) has been made in the theory of plasticity for deter
mining distributions of stress in certain special cases of the partial yielding in which
they do not strictly apply. These are cases in which the material has a definite yield
point and a boundary surface between the elastic region and the yielded portions
develops in the body which under increasing values of the external stresses (e.g., when
a pressure acting on the inner surface of a cylinder increases gradually) moves through
the body. The field of the stresses is gradually displaced within the body under these
circumstances while the stresses satisfy a condition of plasticity. Although in these
cases of partial yielding under variable stresses incremental stress-strain relations of
the form of Eqs. (24-9) or (24-9a) should be used, necessitating a very laborious
integration of the equations with respect to the time t, and Eqs. (24-10) or (24-12) do
not apply, a use of them may be made with considerable advantage. They may
represent averaging stress-strain relations and they supply the simplest, although not
perfectly rigorous, means for obtaining solutions.*

<t>
When the permanent have sufficiently increased, <t>" and
strains
<tt><t>'

become large compared with and approaches the value v" =


v

For monotonously increasing thus Poisson's ratio defined by Eq.


v

(24-1 la) within the plastic zone of body of steel varies continuously
a

with <t>" between the values

£
g

0.3 0.5 .
v

The total dilatation of volume


is

+ = 3(1 -
+

«x «, «. 2i0<frr . (24-14)

and hold but not valid, only the incremental form of the
If

assumptions
1

is
2

3
1

stress-strain relations according to Eq. (24-9a) can be used.


These remarks are made in view of certain recent efforts which have been made to
*

discredit the validity of Eqs. (24-10) to (24-12) in which the negative aspects of theory
were emphasized while positive results furnishing workable expressions of comparable
applicability in special cases are lacking.
SYNTHESIS OF STRAINS 385

Since the permanent strains do not contribute to the dilatation of volume


this must be the same as
u' + fy + (.' = 3far . (24-15)

The condition that the elastic change in volume is proportional to the


mean stress a in the plastic region of a body is therefore also expressed
by the equation
(1
- 2v)4>
= k = const . (24-16)*

Equations (24-12) may be expressed by referring the stress-strain rela


tions to the deviatoral portion of the resultant state of strain. After
subtracting, for example, from tz the mean strain «' = hr corresponding
to the elastic change in volume we have

f. - - e' (1 + v)<t*x
- (3v0 + k)a .

After replacing k here by its expression taken from Eq. (24-16) this
becomes also equal to
e. -e - (1 + r)*(», - a) (24-17)

and we see that the combination of terms (1 + v)<f> with which the shear
ing stresses in Eqs. (24-12) were multiplied appears now also as a factor
in the three equations the strain differences tz — —

t',
expressing

*',
ty
<t>

e, — «'. Furthermore, instead of new function defined by


if

yf/
a

- 2(1
+

w)* (24-18)
*

introduced,1 the resultant strain tensor may be expressed as follows:


is

T». _ ,"
*

v*
'

\
(

'
2
2
41

7«*
^

~
2T'" (24-19)
2

- - "
')

(*.

«.
,

~2
3

in which the left sides constitute six components the devialor


of

the
of

the
resultant strains.
A comparison of the two forms of these stress-strain relations for the
resultant strains [Eqs. (24-12) and (24-19)] with the corresponding two
*
The expression 1/3(1 — 2»)0 = l/3fc may be interpreted as a generalized bulk
modulus, in analogy to the definition of K = E/Z(\ — 2v'). In fact, both expressions
just given are identical since 3k = 1/K.
In the same way that the function which was introduced in Eq. (24-1 la)
4>

the
1

is
<f>'
<t>

sum of two terms: ■» <t>"our new function denned in Eq. (24-18) represents
+

^
</>'

= + f". The constants = \/E and = I/O are equal


<//

a sum of two terms:


\fr'
^'

to the reciprocal values of the moduli of elasticity and rigidity, respectively.


386 THEORY OF FLOW AND FRACTURE OF SOLIDS

groups of the elastic stress-strain relations [Eqs. (24-2) and (24-5)] dis
closes their complete similarity. Instead of the constant factors such as
1/2? and l/G in the first group of equations now the functions 4> and ^
appear. We note that <p and 4/ behave like the reciprocal moduli l/E
1 2(1 + v')
and l/G, respectively, since ~ = - and = 2(1 +
(r a5 ^ v)4>.1

We shall call <p and ^ the flow functions for tension and for shear, respec
tively. and are variable quantities; a material constant.

A;
We

is
4/,
<p,

v
have

= 2(1 + 2v = -- + = 3<t> . (24-20)

k
1

f
^

v)<b

,
,

p
Equations (24-19) can be written in the symbolic form

D. =
\v. (24-21)

in which D, denotes the deviator of the resultant strains.2 They were


expressed in the latter form by A. A. Ilyushin.3
Example Compute Poisson's ratio for a perfectly plastic material after the
1.

yield stress <r0has been reached in tensile test. As long as the tensile stress a si

<
a

the strains are elastic. Let eo = 4>'<robe the elastic strain when the yield point a = <rois
just reached. > «o, <r = <roand the strains will consist partly of an elastic and
If

of
«

a permanent portion, so that the total axial strain


is

- «" - = <t»r°.
+

+
«'

<*>'>o
e

In the mathematical theory of elasticity quantities such as l/E, l/G, l/K


1

(recipro
cal values of the moduli of elasticity) arc called "coefficients of elasticity."

If
instead of the components of the permanent strains those of the tensor of the perma
nent rates of strain are substituted, the corresponding equations for viscous substance
a
<t>

are obtained, which will be considered later. and in the latter case represent
^

the reciprocal values of the "coefficients of viscosity" for tension and for shear,
respectively.
H. Hencky originally expressed these relations in the form
2

[<p

du
- + .p + 3/(m + 1)H dv dw + >p
I

=
I


—Q ~r">
,

+
f

etc-
{"'

di 7+1 Jz Ty
'

~2G~
'
J

u, w denote the components of displacement, m = l/v, variable modulus.


is
v,

(Zur Theorie plastischer Deformationen . . Z. angew. Math. Mechanik, vol.


is 4,
.
,

<p

p. 323, 1924; also Z. Ver. deut. Ing., vol. 69, p. 695, 1925). Hencky's quantity
equal to our 2G>".
<f>,

It believed that our three variable quantities and v introduced above in


iA,
is

the text bring out the analogy in perfect manner between the stress-strain relations
a

for the resultant strains and those which arc valid for an elastic substance, and are
familiar to engineers.
The Relation between the Theory of St. Venant-L6vy-Mises and the Theory of
3

Small Elastico-plastic Deformations, ./. Applied Math. Mechanics Academy


of

Sciences

U.S.S.R. Moscow, vol. p. 207, 1945 (Russian). (Translated by Westinghouse


9,

Research Laboratories.)
SYNTHESIS OF STRAINS 387

Therefore
— >
t — 60
ao

After inserting these values in the expression for v, Eq. (24-lla), we find that Poisson's
ratio

(, (24-22)

as a function of the resultant strains t must in


crease from v {v = 0.3 for most ductile metals)
to v" (»" = 0.5 for moderately large strains
for the metals) according to a branch of an
ordinary hyperbola (Figs. 24-2 and 24-3). It
takes a strain several times the elastic strain «0
at the yield point until v approaches the limiting
value v = \^ of incompressible material. Fig. 24-2. Stress-strain curve for
Example 2. The preceding developments ideally plastic material.
in Example 1 might be applied also to more
general cases of flow, e.g., to ductile metals that strain harden according to a
tensile diagram as shown in Fig. 24-4 in which the curve a = /(e) gradually departs
from an inclined straight line passing through the origin O which in its slope indicates

1 2 3 4 5 6
Fig. 24-3. Poisson's ratio c. Fio. 24-4. Stress-strain curve.

the value of the modulus of elasticity E. Referring to Fig. 24-4, using the angles
a,a',a" under which OP, OQ, and OR are inclined with respect to the strain axis, since

-
<t>'

= cotan a , — — = cotan a cotan a"


,

a a

Poisson's ratio in the transition region of the tensile diagram found to be equal to
is
v

v" — (," -,')—- y" cotano;


cotan a

where a' a constant angle, and v" = are constant values.


\'i
is

v'

This may be further generalized for considering flow of a metal in the strain harden
ing range under a general state of stress whose principal stresses <r1,^2,^3increase with
the permanent strains, provided that the ratios a%/a\, a,/<ri do not change during the
flow.
CHAPTER 25

ISOTROPIC ELASTIC SOLID


The mathematical theory of elasticity is based on the postulates that
the strains are very small, that the components of strain according to
Hooke's generalized law are linear
functions of the components of
stress, that the components of the
displacements in the majority of
cases also remain small quantities
relative to the smallest dimensions
of the bodies and in a few cases may
become comparable in magnitude
with the latter dimensions, and that
the elastic strains are independent
of the manner in which the stresses
have been applied, i.e., independent
of the times during which the latter
have acted. These assumptions
are supported by measurements of
the observed strains or displace
ments within the elastic range in
Fio. 25-1. M. H. Navier (1785-1836). metals, rocks, etc., with some ex
(Courtesy of Pierre Qauja, Institut de ceptions (to the latter belong porous
France, Paris.)
solids, cast metals, etc.), to a suffi
cient degree for most practical applications.
The stress-strain relations for an elastic material were given in Eqs.
(24-2) to (24-8). Let u,v,w denote the small components of the displace
ment v. The components of strain e,, . . . , yv„ . . . are

du dv
=
dw
u
dx ~dz
(25-1)
du dv dv dw dw du
^
.
= y" ~ ~
Yy+dx di + dy~ dx dz

After solving Eqs. (24-4) and (24-5) for the six components of stress

. . , we obtain

388
ISOTROPIC ELASTIC SOLID 389

oz = 2ntz + 3X« , tx, = fiy^ ,

ay = 2/i«„ + 3X« , t„, = /i7»« , } (25-2)


a, = 2nt, + 3X« , t„ = /x7.

In these six expressions, in conventional manner the Lam6 elasticity


constants were introduced : n was written instead of the modulus of shear
G and X for Ev/(l + v)(l — 2v). t denotes the mean strain

U + ty + t.
3

The mean stress a is furthermore

<r* + <r» + c. =
(2M + 3X)« . (25-3)

The preceding equations expressing the law of elasticity must be com


bined with the three conditions of the equilibrium of the stresses:

dx dy dz

fe +
fc + £ + «-o.
X, F,Z are the components of the body forces per unit of volume. After
substituting in Eqs. (25-4) the expressions for the components of stress
using Eqs. (25-2) and (25-1), three partial differential equations for the
displacement components u,v,w are obtained:

flAw + ^ +
lt)-^-
\dx
+ - -)
dy
+
~dz)
+ Z = 0.

Qt Qt Qi
A stands for the Laplacian operator —t + t-j + ^-j- In the absence
of body forces, from Eqs. (25-5)
390 THEORY OF FLOW AND FRACTURE OF SOLIDS

follows. Equations (25-5), named after Navier, are the partial differen
tial equations of the displacement components on which the theory of
elasticity is based.
In the case of a plane distribution of stress (a thin slab stressed in its
plane) without body forces a, = t„ = r,y = 0 we may neglect w. The
preceding equations simplify to the extent that

dx

Ety (25-7)

<?7,„

and the two remaining conditions of equilibrium

^ + ^ = 0, ^ + ^ = 0
(25-8)
dx dy dy dx

are satisfied by assuming that the three components of stress <T,,<ry,T,y are
expressed as the following partial derivatives of a function F(x,y) called
a stress function of Airy :

dW dW ~ dW
*. = =
T"
W' dx~dy'
(25"9)

The equation which must be satisfied by the stress function F(x,y) is


found after substituting the values of the strain components «x,e,,,7n from
Eq. (25-7) in the condition of compatibility of Eq. (25-7) :

This gives

q2
where A now stands f or + .
dx2 dy2
In the case of a plane distribution of stress (<r, = rx, = rn = 0) the
components of stress <r,,avSxy according to Eq. (25-9) are derived from
a stress function F which must satisfy the biharmonic differential equation
(25-11). In a similar way, it may be shown that this is also true in the
case of a state of plane strain = y,, = = 0).
(«,

7x„
The three equations (25-5) may be condensed in one vector equation.
For this purpose, the seven equations (25-2) and (25-3) expressing the
ISOTROPIC ELASTIC SOLID 391

elasticity law and the three conditions of equilibrium (Eq. (25-4)] may be
presented in vector form by single equations. Let the state of stress be
denoted by stress tensor (see Chap. 14)

n = (25-12)

and the state of strain by strain tensor


Try fix
2 2

3pJ
=

<^f
M(W + vT) = t„ (25-13)
y*x y*v

*
2

2
and these two tensors related to each other by using their deviatoral parts
only. The latter are obtained by subtracting from n the state of pure
triaxial equal tensions equal to the mean stress a or the dyadic Lr (where

I
is the idem factor) and similarly by subtracting from strain tensor *
the corresponding pure triaxial extension equal to the mean strain or

<
the dyadic It. These two deviators according to Eq. (24-7) are related
by the condition
D. = 2MZ>,
(25-14)
or
n - Lr = 2/i(* - I«) . (25-15)

Solving for n, noting that the volume elasticity expressed by


is

a = (2/i + 3X)t
we find that
n = 2M4» + 3XIt (25-16)

or after substituting here for * = H(Vv + vV) and for

3t = div v = — + — + —
dx dy dz

that the elasticity law is expressed by Eq. (25-1 or explicitly by


6)

n - n(Vv + vV) XI div


+

(25-17)
v

If S,,Sv)Sr denote the resultant stress in the three plane sections which
intersect each other at one point {x,y,z) and are parallel to the yz,zx,xy
planes, respectively (Fig. 14-6), and = LY + + kZ the body force
is
Y
Q

per unit of volume, stress tensor n also expressed by


is

n = iS, jS„ + kSz


-|-

(25-18)
392 THEORY OF FLOW AND FRACTURE OF SOLIDS

and the condition of the equilibrium of the stresses in an element dx dy dz


[equivalent to the three equations (25-4)] is

or
t f f
+ + + °-° <^"»

V •n + Q = 0. (25-20)

After substituting n from Eq. (25-17) in Eq. (25-20) 1


the differential
equation of the theory of elasticity

M Av + (X + m) grad div v + Q = 0 (25-21)

is obtained.
If the material has no volume elasticity, Poisson's ratio v = J^, X = «o ,
the mean strain « = 0 or div v = 0, the strain tensor is a deviator, the
term Ie in Eq. (25-15) vanishes, and the elasticity law is expressed by

n = 2,i* + la = m(Vv + vV) + la (25-22)

or in components [assuming e = 0 in Eqs. (24-5)]

<r, = 2nt, +*,•••, t„, = M7„, , ' • -


(25-23)

In the absence of body forces the equilibrium condition [Eq. (25-20)] is


V • II = 0 . (25-24)

After substituting n from Eq. (25-22) the differential equation for an


incompressible elastic material is obtained:

n Av + grad a = 0 (25-25)
or in components:

MAM + ^ = 0, MAy +
^
= 0, + 1
- 0. (25-26)
By taking the partial derivatives with respect to x,y,z of the last three

1 In carrying out the scalar multiplications we obtain

V • II = M(V • TV + V -
W) + XV • I div T .

The three terms in this equation are

a* a'
\
(a2

V = div T •W = (div v)V = V (div = grad div v


v)

v

-
,

V-I V .
After inserting these in the expression for V • II, Eq. obtained.
is

(25-21)
ISOTROPIC ELASTIC SOLID 393

equations and adding, we see that the mean stress a must satisfy Laplace's
equation
Ag = 0
(25-27)

determining the unknown scalar function a in Eqs. (25-25) or (25-26).

The mathematical theory of the isotropic elastic substance in the range of infini
tesimal strains has thus been characterized in general terms. It may be in order to
point out at least with a few words that certain materials behaving in a brittle manner
or having a porous structure possessing soft or weak inclusions (cast iron, concrete)
do not follow the linear stress-strain relations expressed by Eqs. (25-2), (25-3), or
(25-14). The stress-strain curve for simple tension or compression for such materials
upon loading can be represented only by a curve <r, = /(«*) within the range of small
strains, and upon unloading a different curve is observed with a much steeper slope
daj/dt,. These materials usually have much elastic hysteresis and show formation of
loops for alternating cycles of loading and unloading (Chap. 3). Attempts to extend
the mathematical treatment to such materials have been variously made. Since these
substances show a distinct volume elasticity, it may in certain applications be suffi
cient to assume for them a linear relation connecting the small dilatation of elastic
nature in volume 3e = «, + t„ + t, with the mean stress a = (<r, + <r, + oi)/3:

e - ci<r (25-28)

whereas the linear equations (25-2) certainly lose their validity.


H. Schlechtweg1 on the suggestion of L. Prandtl has investigated the laws of elas
ticity for brittle materials of this kind. He assumed Eq. (25-28) and proposed a
functional relation connecting the deviator of small strains D, and of stress D, in the
form

= ,
cj).
D.
1 — ca<r — c «ro + cwoV

where ci, . . . , c( should denote material constants and t0 the octahedral shearing

- - - O2
stress:
to ~ H <r,)2 + (<r„ + (<r. + 6(r„, + t„2 + tm') (25-30)

which is proportional to the square root of the second invariant of the stress tensor n.
(The last term c2To'o may become significant when the mean stress a may have large
negative values but may be omitted entirely when a > 0 and brittle materials weak in
tension are to be considered.) The material constants may be adjusted to fit the
observed form of the stress-strain curve for uniaxial tension and compression. Accord
ing to Eq. (25-29) this is supposed to be a common curve passing through «, = 0,

1 Uber ein allgemeines Elastizitaetsgesetz sproeder Koerper, Z. angew. Math.


Mechanik, vol. 14, No. 1, February, 1934. He utilized the elasticity law expressed
by Eq. (25-29) in an interesting application, namely, for a numerical computation of
the distribution of the stresses in a grinding wheel of constant thickness revolving at
high velocity and for predicting the safe speeds of such disks made of brittle sub
stances. He verified Eq. (25-29) on hollow cylindrical cast-iron specimens tested
under various states of combined stress. See his paper on the revolving disk in
Archiv Ing. Wesens, vol. 5, p. 7, vol. 4, p. 119, 1933; and Arch. Eisenhuttenw., vol. 6, p.
507, 1932-1933.
394 THEORY OF FLOW AND FRACTURE OF SOLIDS

vx = 0 having much steeper slopes on the compression than on the tension side of the
diagram, and does not satisfy the rule of antisymmetry of an odd function t, = ip(c),
«, = —<p(—<Ti) usually assumed for plastic materials.
Functional relations connecting the first and second invariants of the strain and
stress tensors may prove useful for extending the elasticity law to the more complex
cases of variable elasticity in solids.

BIBLIOGRAPHY
Fluoge, W. : "Statik und Dynamik der Schalen," Verlag Julius Springer, Berlin.
Frocht, Max M.: "Photoelasticity," Vol. 1, 1941; Vol. 2, 1948. John Wiley & Sons,
Inc., New York.
Geigeb, H., and Carl Scheel: "Mechanik der elastischen Koerper," Vol. 6, "Hand-
buch der Physik," Verlag Julius Springer, Berlin, 1928.
Hetenyi, M.: "Beams on Elastic Foundation," University of Michigan Press, Ann
Arbor, 1946.
L'Hermite, R. : "L'Experience et les theories nouvelles en resistance des matdriaux,"
2d ed., Dunod, Paris, 1945.
Love, A. E.: "A Treatise on the Mathematical Theory of Elasticity," 4th ed., Cam
bridge University Press, London, 1927.
Marcus, H.: "Die Theorie elastischer Gewebe," Verlag Julius Springer, Berlin, 1924.
Nadai, A.: "Die elastischen Flatten," Verlag Julius Springer, Berlin, 1925.
Neuber, H.: " Kerbspannungslehre," Verlag Julius Springer, Berlin, 1937.
Sokolnikoff, I. S.: "Mathematical Theory of Elasticity," McGraw-Hill Book Com
pany, Inc., New York, 1946.
Southwell, R. V.: "An Introduction to the Theory of Elasticity," Oxford Univer
sity Press, New York, 1936.
Timoshenko, S.: "Theory of Elasticity," 1934; "Theory of Elastic Stability," 1936;
"Theory of Plates and Shells," 1940, McGraw-Hill Book Company, Inc., New
York.

'
CHAPTER 26

STEADY SLOW FLOW OF VERY VISCOUS SUBSTANCE


The pressure in a viscous fluid enclosed in a vessel and at rest under
the action of a system of body forces distributes itself according to the
laws of hydrostatics of a perfect fluid. When a very viscous substance
is in steady slow motion, however, shearing stresses are also generated in
it. In many practical applications the distortion proceeds so slowly
that it is permissible to neglect the terms in the dynamic equations con
taining the accelerations of material elements by assuming that the
stresses and body forces belong to a system in equilibrium.
A viscous substance is characterized by the property that the velocities
with which adjacent thin layers of material slide relative to each other
are proportional to the shearing stresses that are generated through this
motion and to the distance between layers. This is expressed by assum
ing that the shearing stress r is equal to
dy
T =
fldi
[see Eq. (2-1)] where dy/dt denotes the corresponding relative rate of
shear and n is a material constant called the coefficient of viscosity.
Suppose that a prismatic bar of such material is loaded in axial direction
by a constant load, e.g., that a glass rod in vertical position is observed
at some slightly elevated temperature with its upper end fixed while a
weight has been fastened to its lower end. The rod will be seen to stretch
very slowly permanently with a constant velocity which is proportional
to the attached load. We do not wash to consider the entire finite dis
tortion of the bar over an indefinitely long period of time t, but only the
small changes of shape of its elements in comparatively shorter times.
Let the x axis be parallel to the axis of the bar and the normal stress under
which it is stretched be <r,. Let be the components of displacement,
u,v,w the velocity components in a point x,y,z at time t, assuming that
£
= i/=f=0at the time t = 0. The permanent unit strains t,,ty,t,
will increase proportionally with the time t with constant rates de,/dt,
dtv/dt, dt,/dt. This motion in the axial direction is expressed by

( = cxt, u = = cx, = = ct, = = c.


^
<*
yx ^ ^ (26-1)
395
396 THEORY OF FLOW AND FRACTURE OF SOLIDS

Furthermore, the stress <r, for a viscous material must be proportional


to the rate of strain

", =
ipTt=,pTx- (26-2)

The proportionality factor <p may be called the coefficient of viscosity


for tension.
If the volume of any small element dx dy dz will not change during
the distortion, the lateral strains must be t„ = t, = —e,/2. An incom
pressible material will stretch with the small constant rates of strain:

= '* = ~
t, =
at
=
7 ' * %
'
(26^;

Generalizing these observations we see that, if a prismatic element


dx dy dz of incompressible viscous material is distorted simultaneously
under the normal stresses <t,,<ty,<i, and shearing stresses Ty,,r,,,r,t, the rates
of strain and shear must be equal to

du 1 / <r„ + <r,\

y"' ~

dv
+
dw -
~
1
Ty'
(26-4)
d~z

'

dy~
m

Furthermore
..... +
du
te+Yy+te^0'
dv dw
+

(26"5)

which the familiar "continuity" equation of hydrodynamics.


is

These six stress-rate-of-strain equations have form similar to the


a

stress-strain relations [Eq. (24-2)] for an incompressible elastic material.


If the quantities appearing in the latter relations: t,,ty,u,yv,,y,,,7,y,E,G
are replaced in the same order by the quantities: i,,iy,i,,yyt,ycx,y,y,<f>,h,
assuming that for the elastic substance Poisson's ratio =
M and pro
c

vided that u,v,w designate the components displacement for the elastic
of

material and the velocity components for the viscous material, one group
of equations transforms into the other group. Furthermore, since for
incompressible elastic material = 2(1 + v)G = 3G, for a viscous
E

= 3m i.e., the coefficient of viscosity for tension equal to


<p

material,
is

three times the coefficient of viscosity for shear.1

This also easily directly verified by computing = 3e,/2 = Tmnx/n for a pure
1

is

extension under tensile normal stress a,. But a, = 2rmu, thus a, = 3/i«, = <pi,
a

and = 3/i.
ip
THEORY OF VISCOUS SUBSTANCE 397

Because of Eq. (26-5) the rate-of-strain tensor is a deviator and the


stress-rate-of -strain relations [Eq. (26-4)] can also be written as follows:

i, =
1
(». -"), 7i»
2
2m 2m'
=
1
(<r, -°), 7«
2
T.,
(26-6)

i, =
1
(». -*), 2
2m 2m'
or more briefly:
D, = 2nD: (26-7)

where D. denotes the deviator of stress, D, the tensor of the rates of strain,
and a the mean stress (a, + <ry + <r,)/3.
The seven equations (26-5) and (26-6), in the absence of body forces,
have to be combined with the conditions of equilibrium [Eqs. (25-4)].
Because of the analogy of the group of equations expressing the elasticity
law (for incompressible material) and the viscosity law, we need not
repeat this computation and can state the results at once. The three
velocity components u,v,w in a viscous material must satisfy the partial
differential equations [see Eqs. (25-26)]

MA« + ox f = 0, „Av + ^-
ay
= 0, nAw + ^
oz
= 0
(26-8)

and the mean stress a must be a solution of the equation

Aa = 0 (26-9)

whereA = —2 + —
+^-2-
Summing up, we see that the theories of the creeping motion of a very
viscous solid and of an elastic material, when both are assumed as incom
pressible, are identically the same. For the former u,v,w denote the small
components of velocity, for the latter those of displacement. The mathe
matical expressions for the solutions satisfying the same boundary
conditions must in fact be the same for both substances if for the coeffi
cient of viscosity n the modulus of shear G is substituted.1
1 As long as Poisson's ratio does not enter into the solution, this may be true also
for certain solutions valid for a compressible elastic solid; e.g., the solutions for the
bending or for torsion of prismatic bars are the same for compressible elastic and for a
viscous substance, etc. Take, for example, the deflection of an elastic cantilever beam
loaded by a force P on its end
PV
y
3EI
(Z length of beam, / inertia moment). It follows that the distance by which the loaded
398 THEORY OF FLOW AND FRACTURE OF SOLIDS

Plane Problem. When the substance moves in planes parallel to the x, y plane
and the velocity component in the direction of the z axis vanishes, w = 0, i, = 0, from
the third of Eqs. (26-6)

a. = = <r (26-10)

and

iz - -iy _ "-2— 13. f m Tu (26-11)

where
du . dv . du dv
--ax' '" =
ay'
y*" =
Ty
+ ax- (2b"12)

The equations (26-8) and (26-9)

MA«+^
= 0, a. A» +
^
- 0 (26-13)

and
A<r = 0 (26-14)

where A = 1 , in the case of a plane flow are solved by introducing a stream


dx* dy*

function. The continuity equation (26-5) on account of i, = 0 is

Thus a stream function tp exists so that

«=g, „--»- (26-16)

Instead of operating with the three equations (26-13) and (26-14) we may eliminate
the mean stress a from Eqs. (26-13) by computing the rotation of the resultant

^
velocity vector which is

•-£-£-*-
After differentiating the first of Eqs. (26-13) with respect to y, the second with
respectto x, and subtracting, we note that the rotation must satisfy the equation

/a Ah d Av\ /du Bv\ _ ,„„

end of a cantilever beam of a viscous material descends uniformly per unit of time or
the velocity of sagging of its free end must be equal to
PI1

where <pis the coefficient of viscosity for tension (<p = 3*<). See also the paper by T.
N. Goodier, Slow Viscous Flow and Elastic Deformation, Phil. Mag., vol. 22, p. 678,
1936.
THEORY OF VISCOUS SUBSTANCE 399

Thus after substituting here the value of &> from Eq. (26-17) a linear differential
equation of fourth order for the stream function

.
AA* 0 (26-19)

is obtained. Solutions of Eq. (26-19) will be used when in the boundary conditions
the velocity components u and v are prescribed.
On the other hand, one may proceed as in Chap. 25, dealing with one of the two
plane problems of an elastic material, by introducing a stress function F, by means of
which the stresses are expressed by the formulas

=
T»=
°'=^' "2
ax'~
'
~din (26"20)

satisfying the equilibrium conditions

°g
dx
+ *£'
dy
= o ,
By
+ fc»
dx
-0 (26-21)

and one determines F from the compatibility condition of the three equations (26-12)

lyi +
tei
- ax-dy
(26"22)

which gives for the stress function


AAF = 0 (26-23)

the same equation as previously found in Chap. 25 [see Eq. (25-11)]. The functions
F and ^ are connected by the relation

- a, = in—
. au
or —
a*F —
a'F
— r
dx'
. ay,
= 4m — -s- • (20-24)
dx dy' dx ay

The following vector equations corresponding to Eqs. (25-22), (25-24), and (25-25)
express the viscosity law
n = ?(Vv + W) + la (26-25)

and the equilibrium of the stresses (in the absence of body forces) :

V •H -0 . (26-26)

Combined they lead to the vector differential equation for extremely slow steady
flow of a very viscous substance :
M Av + grad <r
= 0
(26-27)
and to the scalar equation:
Aj = 0 . (26-28)
v is the velocity vector.1

1 These
two linear partial differential equations arc obtained from the Navier-
Stokes equation expressing the flow of a viscous fluid in hydrodynamics:

?1 + v VV = — grad <r + ^ Av (26-29)


at p p

where p is the density, and from the continuity equation if the two terms on the left
side of Eq. (26-29) containing the acceleration in the fluid are neglected. The pres-
400 THEORY OF FLOW AND FRACTURE OF SOLIDS

Several applications of these equations are discussed in subsequent chapters of


Vol. 2.

BIBLIOGRAPHY
Auerbach-Hoet: "Hydrodynamik," Vol. 5, "Handbuch der Physikalischen und
Technischen Mechanik," J. A. Barth, Leipzig, 1931.
Durand, W. F.: "Aerodynamics Theory," Vols. I, II, III, Verlag Julius Springer,
Berlin, 1935 (Perfect fluids, by T. v. Karman in Vol. II; Viscous fluids, by L.
Prandtl in Vol. III).
Geiger, H., and Karl Scheel: "Mechanik der flussigen und gasformigen Korper,"
Vol. 7, "Handbuch der Physik," Verlag Julius Springer, Berlin, 1927.
Lamb, Sir Horace: "Hydrodynamics," 6th rev. ed., Dover Publications, New York,
1945.
Muller, W.: "Einfuhrung in die Theorie der zahcn Fliissigkeiten," Akadcmische
Verlagsgesellschaft m.b.H., Leipzig, 1932.
Prandtl, Ludwig: Ftthrer durch die Stromungslehre," 3d ed., Friedrich Vieweg A
Sohn, Braunschweig, 1949.
Prandtl, L., and O. Tietjens: "Fundamentals of Hydro- and Aerodynamics," and
"Applied Hydro- and Aerodynamics," McGraw-Hill Book Company, Inc., New
York, 1934.

ence of the "convective" nonlinear term v • Vv in Eq. (26-29) causes difficulties for
finding exact solutions of Eq. (26-29) which are eliminated when this term can be
neglected. For nonsteady motion of a very heavy viscous material the term dv/dt
may be retained.

v
CHAPTER 27

IDEALLY PLASTIC SUBSTANCE


Neglecting the elastic portions of the strains, considering very small
permanent strains and no change in volume during straining (v = we
assume the stress-strain relations1 in the form given in Eqs. (24-10) :

- H(*, + *.)]
- H(?. + ».)] yt, —
, 3<^T|/i ,
y-
- W?. ,
= 3<jrr,, , (27-1)
f, = <j>[*, + *,)] , y» — 3<t>T,v -

These relations may be expressed also by

u = -" = S
<r)

vx
7y (<r, T,
,
2

- - = t„
^
<0

«, , (27-2)
,

)
I

**
=
"2

'
2

where and = 3<£ denote the flow functions for tension and for shear
^

and a the mean stress (<r, + <rv + <r,)/3.


The six strain components on the left sides of Eqs. (27-2) are com
ponents of deviator D, and the six quantities <r, — <T, • ry,, • • •


a

are the components of the deviator of stress D„ so that Eqs. (27-2) in


abbreviated form are also expressed by

D< = |l1,. (27-3)

The six components of strain t,, «„, t„ yyt, y„, y,y are restricted through
the condition that they must be derivable from the three components
u,v,w of the small displacements of the points of the body:


du dv _ dw

dv dw dw . du du dv ,__
,

It understood that for steady motion or flow instead of the strain rates dt,/dt,
1

is

. . the strains*,,... could be substituted in Eqs. (27-1).


.

401
402 THEORY OF FLOW AND FRACTURE OF SOLIDS

The components of stress a,, . . . , t„,, . . . must satisfy three condi


tions of equilibrium [see Eqs. (25-4) assuming XYZ = 0] of the form

ox dy dz

In problems in which displacement components need not be considered,


either for the reason that the boundary conditions are of a static nature
prescribing only components of stress or because one is not interested in
the distortion of the body, one may eliminate the components of displace
ment u,v,w from Eqs. (27-4) by differentiation. Three of the six condi
tions of compatibility of Eqs. (27-4) which can be obtained from the
latter by differentiation should be sufficient conditions to be substituted
for Eqs. (27-4).
The six stress-strain equations (27-1), the three equilibrium conditions
(27-5), and the three compatibility conditions are 12 independent equa
tions which are available for the determination of 13 unknowns, namely,
six stress and six strain components and the flow function <p. One more
equation is needed. The thirteenth equation is furnished by the plastic
ity condition.
Suppose that we introduce the octahedral components of stress and
strain.1 The normal stress on the octahedral planes is the mean stress
a = (a, + <rv + <Ti)/3 and the normal strain is « = (e, + t„ + t,) /3 = 0.
Let <ri,<rj,<r3 and «i,ei,ea be the principal stresses and strains. An ideally
plastic material is defined by the property that the octahedral shearing
stress T0 remains constant* during flow. This is expressed by

I V(*i - «iY~+
to = (*2
- <r3)2 + {<r3
- <n)2
= = const
(27-6)

where a0 is the yield stress in tension.


It was stated previously that r0 may be considered as a measure of the
intensity of stress causing a plastic distortion and that the octahedral
unit shear

7o =
I V(«i
~ + («2
- + («i
- *.)2 (27-7)

is a measure of the intensity of the plastic distortion suffered by the


material. An ideally plastic material flows whenever the stress t0 reaches
its limiting value independently of the shears 70 that are produced.
Any component of unit shear 7 according to Chap. 24 is equal to
y = where t is the corresponding stress. The flow function
1 Papers by author in J. Applied Phys., vol. 8, March, 1937, p. 205; June, 1937,
p. 418.
2 See Eq. (10-27).
IDEALLY PLASTIC SUBSTANCE 403

* ~
3
- ~
(27-8)
3T„ Vf7„
is therefore proportional to the intensity of plastic distortion 70.
If the principal stresses are not known the condition of plasticity
Eq. (27-6) is expressed1 by

to =
3
1
- °.Y + for - 0S + fo - + 6(lV + ry* + r„2)

= (27-9)
V|£2.
The strains «*,...,7„,, . . . may be eliminated entirely after sub
stituting the six expressions for t,, . . . , yv„ . . . from Eqs. (27-2) in
the three compatibility conditions of Eqs. (27-4) of the form

+ ' ^10>
dy* dx* dx dy

Thus three equations will be obtained containing the six components


of stress <r,, . . . , ry,, . . . and the function ^ which together with the
three equilibrium conditions [Eq. (27-5)] and the plasticity condition
[Eq. (27-9)] represent seven equations for the seven unknowns: <r,, ay, a„
tv„ t„, t^, and ^. Equations (27-2), (27-4), (27-5), and (27-9) coincide
essentially with those which were proposed in equivalent form by R. von
Mises.2
We shall not attempt to express explicitly the equations for the general
case of three-dimensional slow steady flow of an ideally plastic substance
for the reason that no hopes exist for finding general solutions of these
equations, but we shall discuss preferably applications of these equations
in special cases such as, for example, rotationally symmetric or two-
dimensional plane distributions of stress in subsequent chapters. With
the introduction of the basic equations (27-1) [or (27-2)] it was presumed
that the components of stress have values in any element of material
1 See Eq. (10-27c). The corresponding equation for 70 was given in Eq. (11-1 la).
' Mechanik der festen Koerper im plastisch-deformablen Zustand. Nachr. kgl. Ges.
Wiss. Gdltingen. Math.-physik. Klasse, 1913. Instead of using the six relations in
Eqs. (27-1) or (27-2) connecting the components of stress with those of infinitesimal
permanent strain, v. Mises expressed the corresponding relations in terms of the rates
of strain i„ e„, . . . which were substituted on the left sides for the strains t,, e„ . . .
in accordance with the way in which these relations had been introduced by de St.
Venant in 1871. Instead of the three conditions of equilibrium [Eqs. (27-5)] v. Mises
introduced the terms expressing the components of the accelerations of material ele
ments in the dynamic equations of motion.
404 THEORY' OF FLOW AND FRACTURE OF SOLIDS

which do not change during the infinitesimal distortion suffered by the


plastic mass. The stress field is assumed to be at rest in the body.1

BIBLIOGRAPHY
Auerbach-Hort: "Handbuch der Physikalischen und Technischen Mechanik,"
Vol. 4 (first part, 636 pp., Elasticity and Plasticity; second part 614 pp., Strength
of Materials), J. A. Barth, Leipzig, 1931.
Bingham, E. C: "Fluidity and Plasticity," McGraw-Hill Book Company, Inc., New
York, 1922, 440 pp.
"First and Second Report on Viscosity and Plasticity," prepared by the Academy of
Sciences, Amsterdam, Vol. 1, 256 pp., 1935; Vol. 2, 287 pp., 1938.
Geiger, H., and Karl Scheel: "Handbuch der Physik," article on plasticity, Chap.
6, pp. 428-500, Vol. 6, Verlag Julius Springer, Berlin, 1928, 632 pp.
Geiringer, Hilda: "Fondements mathematiques de la theorie des corps plastiques
isotropes," Memorial des sciences mathematiques, Gauthiers-Villars & Cie, Paris,
1937, 89 pp.
Hodge, P. G.: "An Introduction to the Mathematical Theory of Perfectly Plastic
Solids," Graduate division of Applied Mathematics, Brown University, Provi
dence, R. I., February, 1950, 396 typewritten pages.
Houwink, R. : "Elasticity, Plasticity and Structure of Matter," Cambridge Univer
sity Press, London, 1937, 376 pp.
Ilyushin, A. A.: "Plasticity" (Plastichnost), Moscow and Leningrad (in Russian),
1948.
Kachanov, L. N., N. M. Belaiev, A. A. Ilyushin, W. Mostow, and A. N. Gleyzal:
"Plastic Deformation, Principles and Theories," Mapleton House, Brooklyn,
1948, 192 pp.
Prager, W.: "Mecanique des solides isotropes au dela du domaine elastique,"
Memorial des sciences mathematiques, Gauthiers-Villars & Cie, Paris, 1937.
See also a number of his papers on the mathematical theory of plasticity in reports
published by the Graduate Division of Applied Mathematics, Brown University,
Providence, R.I., 1947-1948.
Reiner, M.: "Ten Lectures on Theoretical Rheology," Rubin Mass, Jerusalem, 1943,
164 pp.
Sokolovsky, W. W.: "Theory of Plasticity" (in Russian), published in Moscow by
the Academy of Sciences of the U.S.S.R., 1946, 306 pp.
van Iterson, F. K. J.: "Plasticity in Engineering," Blackie & Son, Ltd., Glasgow,
1947, 174 pp.

1 Thus the cases of partial yielding which may arise in bodies whose material has a

well-defined yield point and an elastic range of strains are a priori excluded. In the
latter cases the elastic portions of the strains which were neglected according to the
assumptions made must be considered and when a plastic zone has developed in the
body the stress field in general will not remain at rest but will sweep through the body
with the advancing plastic front, making it necessary to postulate incremental stress-
strain relations instead of the ones expressed in finite form in Eqs. (27-1).
CHAPTER 28

FURTHER SIMPLE AND COMPOSITE PLASTIC


AND VISCOUS SOLIDS

28-1. Constant Maximum Shearing Stress. A forerunner of the


theory of an ideally plastic substance treated in the preceding chapter
was the earlier theory of the plane two-dimensional flow of a plastic
material developed during the eighteen seventies by the great French
engineer Barr6 de St. Venant.1 It differed essentially only in one respect
from the present theory of an ideally plastic substance, namely, in the
form of the condition of plasticity. To illustrate the difference between
the two forms of the condition of flow and for comparison with some
further related theories of flow of more recent origin, the shapes of the
corresponding surfaces of yielding will be compared. As stated in Chap.
15 the condition of plasticity can be represented by a surface in a rec
tangular system in which the principal stresses are the coordinates. For
an ideally plastic substance the surface of yielding [Eq. (15-16)] is a
circular cylinder with an axis inclined at equal angles with respect to the
a\jTi,at axes.
Tresca2 around 1865 concluded from his tests with lead extruded under
high pressures that this metal flows when the maximum shearing stress
Tumi reaches a certain value. According to this hypothesis, a material
flows whenever either of the differences of the principal stresses reaches a
constant value [Eq. (15-20)]:
o~i
— <rt
= +oo , 02

<r»
= ±<ro , fft
— (ri = +<7o . (28-1)
These six equations represent three pairs of parallel planes through which
a regular hexagonal prism is inscribed in the cylinder defined by Eq.
(15-16). Based on Tresca's interpretation of the latter's extrusion tests,
St. Venant developed his theory for the plane flow of a plastic material
which will be treated in Chap. 37.
1 Compt. rend., vol. 70, p. 368, 473, 1870; vol. 73, p. 86, 1098, 1181, 1871; vol. 74,
p. 1009, 1083, 1872. For a critical review of these early attempts for establishing a
theory of plasticity, cf. Todhunter and Pearson, "History of the Elasticity and
Strength of Materials," Cambridge, 1886. Also Brillouin, Marcel, Ann. phys.,
vol. 13, pp. 75-113, 1920; Praoer, M. W., "Mecanique des solides isotropes au dela du
domaine elastique," Vol. 87, Memorial des sciences mathematiques, Gauthier-Villars
& Cie, Paris, 1937, 66 pp.
1 Compt. rend., savants etrangers, 1865, 1868, 1870, Paris.
405
406 THEORY OF FLOW AND FRACTURE OF SOLIDS

28-2. Generalized Ideally Plastic Substance. The ideally plastic


substance was defined in Chap. 27 by assuming that at the limit of plas
ticity the octahedral shearing stress t0 has a constant value to = y/2 <r0/3
((To This substance yields under a state of simple
yield stress in tension).
compression (d =
o%
= 0, 0-3 = — tr0) at the same absolute value of the
stress as under simple tension (o-i = o~o, <ri = <r» = 0). This is true also for
a material which yields when the maximum shearing stress r,., reaches a
constant value r^, = o-0/2 (Sec. 28-1). We mentioned in Chap. 15, when
the theories of strength were reviewed, however, that both these condi
tions of plasticity fail to represent cases of flow of materials in which
the yield stress for simple tension is different from that for simple com
pression. When the limiting states of stress depend on the mean stress
ffm = (<ri + <?z + c3)/3 one may, according to 0. Mohr, assume, for exam

ple, that the limiting value of the shearing stress t causing a plastic dis
tortion is a function of the normal stress <r in the plane sections in the
material near which the permanent slip originates in thin layers. As
mentioned in Chap. 15 this amounts to assuming that at the limit of
plasticity the difference of the major and minor principal stresses a\ — <r3
is a function of their sum <ri + <r3 (ai > <r2 > 0-3) and that the major prin
cipal stress circles in the <t,t plane graphical representation of the limiting
states of stress possess a common enveloping curve. The latter curve
therefore touches also the two principal stress circles of unequal radii
representing the states of simple tension and compression. The general
Mohr envelope of these major stress circles contains the special case in
which all these circles have equal radii, i.e., in which this envelope con
sists of two straight lines parallel to the a axis in the a,r plane. (This
case was treated in Sec. 28-1, when r^, = o-0/2.) However, the major
principal stress circles may also have two straight lines for their envelope
inclined at equal angles with respect to the a axis, which represents the
plasticity condition by J. Guest:1

<ti

(73
= Ci + c,(<ri + «r3) . (28-2)

Postulating the existence of a Mohr envelope the equation

ffi - <r3
= /(<rt + <r3) (28-3)

expresses the condition that the radii {n — <r3)/2 of the major principal

1 The linear equation <ri — at = C\ + c«(<ri + a 3) as a limiting condition of failure


in shear has been proposed by the French physicist 0. A. Coulomb (Mim. math, phys.,
vol. 7, 1773); it was also used by C. Duguet (Limite d'clasticit<5 et resistance a. la
rupture," Paris, 1885.) The limiting condition of slip in the plane problem of the
equilibrium of a loose material such as sand coincides with this condition when d — 0
and the stresses are compression stresses. See Sec. 15-130.
SIMPLE AND COMPOSITE SOLIDS 407

stress circles at the limit of plasticity are a function of the distances of


their centers (cri + <ra)/2 from the origin of the <t,t system of rectangular
coordinates. We note that in this plasticity condition the intermediate
principal stress <n does not appear.
When O. Mohr established Eq. (28-3) he was concerned with formulat
ing a condition of failure by slip (and fracture) in stressed materials. As
L. Prandtl1 remarked, one can generalize the concept of a perfectly
plastic substance (not having the property to strain harden), however,
by assuming that the material elements in a body continue to deform
permanently indefinitely when Eq. (28-3) is satisfied. We are thus led
to cases of slow flow in solids in which the yield stresses depend also on
the mean stress am and may think of combining Eq. (28-3) with the three
conditions of equilibrium and with some form of the stress-plastic-strain
relations. For the latter we may, for example, assume since Eq. (28-3)
does not depend on the intermediate principal stress <r2, that whenever
o-i t^ <ri or 0-2 y* (T3 material elements do not deform in the direction parallel
to <T2 so that, if f i > tr2 > o-3 the principal strain ej = 0 and when either
o\ = <T2 or when o-j = <r3 the principal permanent strains «i = «a or «j = e3,
respectively. This would lead to the necessity of subdividing the plastic
body in regions, in which the analytic expression of the condition of flow
changes its form according to which of the three principal stresses a\,ai,a%
is the major, the intermediate, or the minor stress or according to whether
regions in the body exist in which two of the principal stresses become
equal to each other. This introduces certain discontinuities in the
analytic form of the flow condition and of the stress-strain relations.
These discontinuities also become apparent when the corresponding
surface of yielding fi(<r^at,<Xi) = 0 is considered. Since Eq. (28-3) does
not contain the variable <n, it represents in the system of rectangular
coordinates a\,oi,os a surface normal to the <j-i,er3 plane. If the coordinates
<n,<T2,ff3 are interchanged in cyclic order, three such surfaces are obtained,
intersecting each other. Thus the surface of yielding is represented by
a slender body having six curved faces and edges, as was mentioned in
Chap. 15. Equations relating the two geometric representations of the
generalized condition of flow Eq. (28-3) (treated as an envelope in the
<r,T plane and as surface of yielding in the o\,<n,a% three-dimensional space)

have been discussed by C. Torre2 recently in much detail. He considers,


1Nach. Ges. Wiss. Gotlingen, Malh.-phys. Klasse, 1920, p. 74; Z. angew. Math.
Mechanik, vol. 1, p. 15, 1921 ;Proc. 1st Intern. Congr. Applied Mechanics, Delft, 1926,
p. 43.
• "Ueber den plastischen Koerper von Prandtl. Zur Theorie der Mohrschen Grenz-
kurve." Also "Einfluss der mittleren Hauptnormalspannung auf die Fliess- und
Bruchgrenze," Oesterr. Ing. Arch., vol. 1, pp. 36-50, 316-342, 1948.
408 THEORY OF FWW AND FRACTURE OF SOLIDS

for example, the case in which the envelope of the major principal stress
circles in the o,t plane is an ordinary parabola,

t2 = 2p(<r* - a) (28-4)

(p,<r* are constant values), a case which had been proposed by A. Leon.1
In terms of the principal stresses ci and at this is equivalent to assuming
a flow condition

(<n
- <r3)2 + 4p(<r, + <r3)
= 4p(2cr* - p) . (28-5)

Using Eqs. (28-4) and (28-5), Torre worked out the expressions for the

Fio. 28-1. Slip lines in section of a heavy-walled cylinder stressed by internal pressure
{after C. Torre) when the condition of plasticity depends also on the mean stress.

distribution of the stresses in a heavy-walled cylinder stressed by an


internal pressure (Fig. 28-1) and in the plane problem of the flow in a
blunt wedge, a case which was previously worked out by Prandtl by using
Eq. (28-1).
When discussing the surfaces of slip we mentioned (in Sec. 15-13/) that
in materials in which the yield stress depends on the mean stress an the
angle under which the surfaces of slip are inclined with respect to the
major principal compression stress changes with the states of stress whose
major principal stress circles are tangent to the envelope Eq. (28-3).
This is illustrated by the curves of slip reproduced in Fig. 28-1 for the
case of the flow in a heavy-walled cylinder yielding according to the con
dition of Eq. (28-5) and to Torre's computation. Whereas these curves
consist of two families of orthogonal logarithmic spirals (see the photo
graph Fig. 37-12 and Fig. 37-13) for a ductile metal yielding under the

Ing. Arch., Vol. 4, p. 421, 1933; Bauingenieur,


1
vol. 15, p. 318, 1934. Further
papers are quoted by C. Torre, loc. cit.

"v
SIMPLE AND COMPOSITE SOLIDS 409

condition T„t = const, the system of slip lines in Fig. 28-1 deviates quite
considerably from an orthogonal pattern of trajectories.
Some of the mathematical difficulties inherent in conditions of plasticity
in which one of the principal stresses does not appear [as in Eqs. (28-3)
or (28-5)] which were mentioned above can be eliminated if one introduces
instead of the condition Eq. (28-3) a mathematically related condition
of flow in which the octahedral shearing stress t0 is a function of the mean
stress am = {a i + <r2 + <r3)/3,
to = /(»„) . (28-6)
Equation (28-6) is represented by a surface of revolution having its
axis inclined at equal angles with respect to the axes a 1,,2,<r3. An example
in the shape of a circular cone was mentioned in Sec. 15-13A, Eq. (15-29)
which could be used instead of the Guest condition of flow.1
28-3. Theory of the Strain Hardening of Metals. The true flow stress
in a tensile test of a soft metal at normal temperature when plotted as a
function of some measure of the permanent strains defines the strain hard
ening curve of the metal in tension. Curves of similar character may be
obtained from compression, torsion, or other types of tests made with
metals. Concerning the property of all such curves indicating an increase
of the flow stresses with the permanent strains, the question may be asked
whether a general strain hardening function could be defined connecting
a more general stress and a similar strain variable, expressing the behavior
of a metal, from which the strain hardening curves corresponding to the
simple states of stress (tension, compression, etc.) could be derived.
Attempts for defining a general strain hardening function are old,2 but
certain difficulties obstructed its establishment, among which one was
that new convenient strain variables had to be found for expressing large
permanent distortions in the metals and another was that the metals
become anisotropic gradually in their properties when the permanent
strains have increased to considerable magnitudes. Two working
theories may be proposed, however, one for comparatively small perma
nent strains and the second for moderately large strains. In both of
these the usual assumption of an isotropic flow of incompressible material
is retained. In most cases it is sufficient to neglect the elastic portions
of the strains, since the permanent strains will be predominant even
though they may still remain small quantities, as in the first case which
1 It should be noted that if this surface intersects the
space diagonal a\ = ai = at
a portion of the surface must be excluded around this axis since a material cannot
yield under a state of equal triaxial stresses.
2 P. Litdwik
(loc. cit.) in 1909 introduced a general "flow curve" for the ductile
metals assuming that the maximum shearing stress is a function of the maximum
unit shear.
410 THEORY OF FLOW AND FRACTURE OF SOLIDS

was just mentioned. Furthermore, it is assumed that the mean stress


<r = (<ri + 0-2 + <r3)/3 has no appreciable influence on the magnitude of
the shearing stresses that are required to produce the permanent deforma
tion. Although they were not all stated expressly, the theories of simple
plastic bending and torsion which were treated in several previous chap
ters were based on these assumptions.
Assuming small purely plastic strains satisfying the condition

«x + + ««
= 0
«»

the theory of the ideally plastic substance (Chap. 27) may be generalized
by postulating that the octahedral shearing stress t0 is a slowly, monoto
nously increasing function of the octahedral shear 70 but that t0 does not
depend on the mean stress o\ Suppose that during the straining of the
substance the stresses are increased in such a manner that the ratios of the
principal stresses <ti/<j\ and 0-3/0-1 remain constant. It is understood, if
<rj/ffi and 0-3/0-1 are constant, that instead of the rates of strain the small
components of permanent strain may be substituted in the stress-rate-of-
strain relations and instead of the latter plastic stress-strain relations1
according to Eqs. (27-1) or (27-2), and a general strain hardening func
tion or its inverse function

to = f(yo) or 70 = g(r0) (28-7)


may be introduced, in which t0 is given by Eq. (27-9) and 70 by

- ..)* + -
yjd*

~ eyy
7o =
I + + 7,.J +
+

(«, («. «,)« (7x„2 7.x2)

.
I

(28-8)
The flow functions <p and for tension and for shear are
^

_. To _ 9(ro) _ 7o _ _ To ,„„
,
.

Qs
.

3t0 3t0 3/(7o) To

From a mathematical point of view we shall consider the function t0 — /(>») as


1

the logical generalization either of the condition of flow to = const (t0 independent of
70) of the ideally plastic substance or wc may see in to = f(yo) a generalization of the
linear stress-strain relation t0 = Gy<, valid for an elastic material (which must be
assumed as incompressible the condition tx + «, + t, = to be satisfied). In
if

is
0

the following developments we shall consequently use the term "plastic" in a general
sense for characterizing those materials in which the flow stress depends only on the
permanent strains (but not on the rates with which the latter change). These
materials have variable modulus of deformation l/if*. W. Prager (Strain Harden
a

ing under Combined Stresses, J. Applied Phys., vol. 16, No. 12, pp. 837-840, 1945)
discusses general restrictions as to the mathematical forms expressing the relation
between the deviators of stress and of permanent strain and attempts to»construct
such relations in power developments subject to these restrictive conditions.
SIMPLE AND COMPOSITE SOLIDS 411

The flow function for shear tp is given through the cotangent of the
angle under which a cord drawn from the origin to a point 7o,to of the
general strain hardening curve is inclined with respect to the 70 axis and
can be computed if this curve is experimentally known. If a tensile
stress-strain curve <ra = F{ta) has been observed, the general strain hard
ening function r0 = /(70) is found from the former through the substitu
tions <r0 = 3t0/\/2 and *0 = 7o/\/2 or from

r.,/(T„)„V?f(^. (28.10,

According to the preceding theory the strain hardening function for


simple compression should be expressed by the equation <re = — F{ — to)
if the strain hardening curve for simple tension has the equation

«r«
= F(t0) ,

i.e., they are congruent curves. For metals in which this condition does
not hold and for which the compression curve differs in its ordinates
from the tensile curve for the same absolute values of the strains the
general strain hardening curve may be assumed in the form

70 = H(t0,<to) (28-11)

where 70 denotes the permanent unit shear, r0 and <r0 the shearing stress
and the normal stress [<r0 = (<ri + <n + <r3)/3 equal to the mean stress]
in the octahedral planes of the strain or stress tensors.
In an attempt to extend the strain hardening range of permanent deforma
tion to finite strains of moderate magnitude after assuming that the concept
of isotropic behavior is preserved, the stress-strain relations Eqs. (27-1)
which were valid for an infinitesimal distortion (they contained the
conventional strains) may be expressed now in differential form:1

dii = Lri
a*J

d<f>
,

dh = - "-~^\
(at

d<j, (28-12)
>
,

dis = lot -
—J

d<t>
,
2

where dii,dit,dii denote now the increments of the principal natural

All the components of stress such as vi, , n,


1

. . . to appearing in the subse


.

quent equations (28-12) to (28-20) are denoting true stresses, computed for the actual
areas of cross sections under load.
412 THEORY OF FLOW AND FRACTURE OF SOLIDS

strains. The corresponding increments of the principal natural shears


are equal to

0-2
— <T3
dyi = d«2
— dig = 3 t\ =
d<f> , T]
2

dy^ = dis — dii = 3 d<j> ti , ti


T2 = ———
ffl
ff.1
^ , f (28-12a)
2

dji = — diz = 3 — ffl <T2
dii d<t> t3 , TX

ti,ts,t3 are the true principal shearing stresses. Since a similar equation
holds for the increment of the natural octahedral unit shear

dy0 = 3rf0TO (28-126)

we see that the increment of the flow function

d4>
= ~ (28- 12c)
OTo

is known as soon as the strain hardening function of the metal

to = /(7o) (28-13)

has been determined from tests. For computing €i,(2,e3 Eqs. (28-12)
using (28- 12c) must be integrated.
The preceding equations may be integrated and expressed in finite
terms when the principal stresses ffi,ff2,ff3 during the increase of the
external load are changed in such a manner that their ratios:

2
ffi
= k, ,

ffi
- ks (28-14)

are kept constant. (This has been the case, for example, in the tests on
the yielding of copper and mild steel under biaxial stresses by E. A. Davis
which were reported in Chap. 17-7.) According to Eq. (28-14) t0
increases in the same proportion as the principal stress <n; therefore also

to =
I V(ffi
- ff2)2 + (ff2
- ff3)2 + (ffs
- <r.)2
= fen (28-15)

where k denotes the constant

k - | V(l - fc2)2 + W*
- k3y + (k, - l)2 . (28-15a)

The first of Eqs. (28-12) may now be expressed as follows:

,.
dtl = ( _
^
+ o*\ __ _ ( _ _____
___j
ff2

^i
+
J
_
dy»
'
kt fc3\ dy0 fl>o
(28-16)
lr%
SIMPLE AND COMPOSITE SOLIDS 413

and integrated from the unstrained condition of the body to a strained


condition,

Thus, if a new flow function * is defined

' - & - & - m> <»-■«

the three stress-strain relations [Eqs. (28-12)] may be integrated and


expressed in finite form as follows:

" = *( + °*\
2~ )'
e, - *- (
<y% + <ri\
^r,

2~ (28-19)

-- =
)'
*^2 2~)>
where denote the natural principal strains.
It to the credit of E. A. Davis that he made use of the latter stress-
is

strain relations in finite form for evaluating the strain hardening function
to = /(7o) from his tests on copper and steel up to distortions of consider
able magnitude which were reported in Chap. 17.
1

If the strain hardening curve rises quite gradually from the origin, as
for some soft annealed metals, the function /(70) in Eq. (28-10) for small
strains may be approximated by assuming that power function,
is
it

To
= T°*
fe)m (°^m^V (28-21)
'

(ro*,7o*,wi are constants). In simpler cases of the flow under two- or


three-dimensional distributions of stress (of a heavy-walled cylinder
under internal pressure or of a thin disk having circular hole, etc.) the
a

If the quantities and „ defined in Chap. 16 are utilized, this amounts to assuming
1

that for finite strains not only


2<r2
— ,1 —
<r> dii — dii — dio
2

(28-20)
o\ — at dii — dii
but that the right side of this equation may be expressed also in the finite form

_ 2ii — —
ii

«i _

ti e5

provided that the conditions of Eq. (28-14) are not violated.


414 THEORY OF FLOW AND FRACTURE OF SOLIDS

corresponding differential equation can be integrated and exact formulas


can be computed for the stresses or strains in metals yielding according
to the strain hardening function [Eq. (28-21)]. For example, in a thin
disk stretched in its plane by a stress a uniformly in all directions and
having a small circular hole r = T\ (Fig. 28-2) a factor of the concentration

Fio. 28-2. Stresses around hole in disk in viscous (m — 1), in creeping (1 > m > 0), and
in plastic (m = 0) substance.

of stress k denned by the ratio of the circumferential stress at at the surface


of the hole (r = n) to the stress a at r = oo can be computed:1

(28-22)

It is assumed hereby that the entire disk yields. This factor & is a func
tion of the exponent m of the power law expressed by Eq. (28-21) which
is very nearly approximated by the linear function

k = 1 + m , (Ogtngl) (28-23)

(see Fig. 28-3).


If
m is taken equal to zero, the concentration factor k becomes unity
and in Eq. (28-21) t0 = const. This is the case of an ideally plastic sub
stance; when the exponent m in the power function is chosen equal to one
we obtain a second limiting case and a stress concentration factor k = 2.

1 /. Applied Phys., vol. 8, p. 428, June, 1937.


SIMPLE AND COMPOSITE SOLIDS 415

<tt>
The flow function in Eq. (28-9) constant and the stress-strain rela

is
tions become linear equations. This second limiting case of the power-
function law coincides with the law of elasticity for incompressible
material and we have = in

2
k
agreement with the theory of elas
ticity. Formula (28-22) becomes ////JV
/
indefinite for m = but one //

if
substitutes in for the exponent S //
it

5
3m) = a, so that when

2m/
(1

m = « = oo, appears in the // / /

f
v

/
form = lim
$ which y
+

s
lim.
k

/P/ s
0

7
/
equal to = 1.396. Equation
is

///s
I.2\

(28-22) as well as (28-23) shows /s//


that the factor of concentration of 1.0.

stress in a disk having a circular EXPONENT m -*■


hole which yields under the strain Fio. 28-3. Factor of stress concentration
hardening law [Eq. (28-21)] varies plotted versus ex-
(r
+ 3m )2m/(l-liw)
between the values = and = m of strain hardening law to
k

ponent
2
1

TO*

corresponding to the two limiting for disk with hole.


cases of the ideally plastic and the
perfectly elastic substances, respectively.1
A. A. Ilyushin2 remarked that in bodies yielding according to the power-
function law Eq. (28-21) the internal stresses become proportional to
their boundary values which represent the loads under which they are
deformed permanently. Alice Winzer and W. Prager3 characterized

The preceding developments may be applied to the same shapes of solids under the
1

states of stress to which reference was just made they slowly deform permanently
if

under constant loads in the steady range of creep at elevated temperatures, provided that
the law according to which these solid bodies creep expressed by the power function
is

--(&)" m
g

1)
(0

(28-24)
£

where 70 now expresses the steady rates of the octahedral unit shears with which the
material deforms under the shearing stresses to (to*, 70*, m are constants). All that
changes in the mathematical expressions that for unit shears in the creep range the
is

steady creep rates dy/dt = are to be substituted. This follows from the develop
7

ments in Sec. 28-4. The upper limiting case for the exponent m = furnishing the
1

factor of concentration of stress — in the case of the creep of disk having a hole
k

a
2

coincides with the steady slow deformation of a perfectly viscous solid (Fig. 28-2).
The Theory of Small Elastic-plastic Deformations, /. Applied Math. Mech.,
'

Moscow, vol. 10, p. 347, 1946, available in English translation at the David Taylor
Model Basin, Navy Department, Washington, D.C.
On the Use of Power Laws in Stress Analysis beyond the Elastic Range, J. Appl.

Mechanics, ASME preprint, December, 1947.


416 THEORY OF FLOW AND FRACTURE OF SOLIDS

these cases by stating that for them the ratio T0/y0 [they called it the
"secant modulus" of the strain hardening function t0 = /(70)] is pro
portional to a power of the octahedral shearing stress to, to which we may

<t>
add that this is also true of the flow function which proportional to

is
the reciprocal value of the "secant modulus."
28-4. Steady Stage of Creep. metallic bar at elevated temperature

A
under constant load in tension stretches continuously slowly or creeps.
a

After the bar loaded stretches first, when the temperature not too
is

is
it
high, with comparatively faster rates, but these rates of creep decrease
soon until second period reached during which the rate of stretching

is
a

remains practically constant over a long period of time (see Fig. 28-4).
Subsequently the rates of creep increase and the bar starts to neck before
finally breaks. At a given temperature
it
this steady or minimum creep rate is only a
the stress. During this second

of
function
stage of creep the behavior of metal may

a
well be compared with that of a viscous
TIMEt material, the essential difference being, how
Fia. 2&4. Creep curve. ever, that while the rate of strain in per

a
fectly viscous substance proportional to the

is
stress this not the case for the creep of the ductile metals. In spite of
is

this difference in behavior we shall list the phenomena of steady creep in


solids among those of generally viscous nature in contrast to the gen
a

erally plastic phenomena which were discussed in Sec. 28-3. The term
"viscous" in this sense characterizes those materials in which the flow
stress depends only on the rate of permanent strain (but not on the strains
themselves). These materials have variable coefficient of viscosity.
a

Neglecting "primary creep" during the first stage the flow during the
second, steady stage of creep may be generally characterized by assuming
that the octahedral shearing stress to a monotonously increasing func
is

tion of the rates of shear 70 = dy0/dt. Assuming small rates of creep


and small creep strains thus the theory of viscous substance (Chap. 26)
a

generalized by assuming that


is

ro = /(70) or iro = S'(to) (28-25)

where the "minimum" octahedral rate of shear to be computed


is

70
from the expression which was given for 70 in Eq. (28-8) after substituting
in the rates of strain iz, «„,... instead of the strains tz, «„,....
it

This true also for the stress-rate-of-strain relations which follow from
is

the stress-strain relations [Eqs. (27-1)] after substituting in them the


ix,

rates of creep «„ . . instead of the strains.


.

'
SIMPLE AND COMPOSITE SOLIDS 417

For the creep function has been proposed1 either a power function

fej
to = t* <

(n
(28-26)

1)
,
or the hyperbolic sine function

7o = 7* sinh UU (28-27)

where t* and 7* are material constants.2


Similarly theory of the flow of
as the strain hardening metal (Sec.

a
28-3) was based on strain hardening function t0 = /(70) which again
a

could be considered as the mathematical generalization of the linear


stress-strain function t0 = G70 of an (incompressible) elastic material, the
theory of the steady creep of solids can be based on creep function

a
To = /(70) as the logical generalization of the linear stress-rate-of-strain
relation to = M7o of the viscous solid. The five fundamental types of
simple solids (including the ideally plastic solid) are represented by the
graphs of these functions in Figs. 28-5 to 28-9. We may add that the
8

The power function [Eq. (28-26)] has been used by several investigators: by R. W.
1

Bailey (Utilization of Creep Test Data in Engineering Design, J.


Inst. Mech. Engrs.,
London, vol. 131, pp. 131, 269, 1935; also Engineering, vol. 140, p. 595, 1935. Bailey
uses an assumption equivalent to Eq. (28-26), but stress-rate-of-strain relations
different from those quoted above in the text); by H. F. Tapsell and A. E. Johnson
(Engineering, vol. 150, p. 24, 1940); by F. K. G. Odquist (Creep under the Action of
Combined Stresses, Ingeniors Vetenskaps Akademien, Stockholm, No. 141, 1936);
and others. The hyperbolic sine function Eq. (28-27) was proposed by L. Prandtl,
(Z. angew. Math. Mechanik, vol. p. 85, 1928), by P. G. McVetty and by the author
8,

(Proc. ASTM, vol. 43, p. 735, 1943; Trans. ASME, Davenport spring meeting, 1943);
also in "Th. v. Karm&n Anniversary Volume," Applied Mechanics, 1941, paper by
a

author, p. 237.
One may interpret the slope dro/dyo as a variable coefficient viscosity it. Accord
2

of

ing to the power function Eq. (28-8) = dro/dyo = « at the origin while for the
11

hyperbolic sine function Eq. (28-27) the slope = t0*/7o* at the origin. finite
11

value of the viscosity at very small rates of shear (microcreep) should be expected
n

from physical considerations. The hyperbolic sine function was deduced by H.


Eybing (Viscosity, Plasticity and Diffusion as Examples of Absolute Reaction
Rates, J. Chem. Phys., Vol. p. 283, 1936) from the so-called chemical rate theory of
4,

statistical mechanics in his work on viscosity and creep and applied by W. Kautzmann
(Metals Technol., vol. 1941) to the creep of copper.
8,

This analogy between the theories of strain hardening and of the steady stage of
3

creep of metals was further discussed in some detail in the author's paper quoted on
p. 414, where various formulas for the distributions of the stresses in thick-walled
cylinders were given for both cases and factors of stress concentration evaluated
around a hole assuming power function for to = /(70) and to = f(yo), respectively,
a

for both types of flow.


418 THEORY OF FLOW AND FRACTURE OF SOLIDS

strains in these five graphs are either purely permanent or purely elastic.
For this reason these five ideal substances are examples of simple solids.
28-5. Plastico-viscous Substance. Recalling that a relation a = /(t) could be
used for expressing the tensile behavior of a ductile metal in the strain hardening
range and a = /(«) in the creep range, i.e., at comparatively low and at high temper
atures, one may assume that the flow stress a of a metal in the intermediate range of
temperatures might depend on both these vari
ables and be expressible by a function of the
strains e and rates of strain i:

(28-2Si

ELASTIC SOLID STRAIN HARDENING By combining the meanings attributed to


METAL
the terms "plastic" and "viscous" in Sees.
Frq. 28-5. Flo. 28-6.
(28-3) and (28-4) materials following the defor-

STRESS <r

VISCOUS SOLID STEADY STAGE OF


CREEP OF METALS
Fio. 28-8.


IDEALLY PLASTIC
MATERIAL
Fio. 28-9. KRMAN STRAIN f
Figs. 28-5 to 28-9. Five types of Flo. 28-10. Block diagram of stress sur
simple ideal solids. face for plastico-viscous substance.

mation rule of Eq. (28-28) may be called "plastico-viscous." Equation (28-28) as a


function connecting three variables: a,t,i can be represented by a surface. « and i may
be used as the independent variables and a as the dependent variable in rectangular
coordinates. Such a "stress surface" a = /(«,e) is shown in Fig. 28-10 for tensile
tests. The curves obtained as the intersections of this surface with planes parallel
to the a,t plane would represent the tensile stress-strain curves of what might be called
the "constant strain rate tests" (i = dt/dt = const) and the curves situated in the
planes « = const would define the dependence of the flow (yield) stress on the rates
of strain.1 Whether such a stress surface exists for a given metal at a given temper
ature is not a priori clear. All that can perhaps be said is that, if « and < are changed
by small increments, the flow stress a will increase by the increment

, da
— , da ,.
da =
dt
dt +—
dt
dt . (28-29)

da/dt measures the rate of strain hardening in a constant rate test and da/di may be
interpreted as a variable coefficient of viscosity. Assuming that these two quantities
1 Some
properties of these surfaces were described in Trans. ASME, vol. 55, 1933
(APM-55-10, p. 61), and in J.
Applied Mechanics, March, 1936. The surface as
represented in Fig. 28-10 passes through both the i and • axes.
SIMPLE AND COMPOSITE SOLIDS 419

ilepend on the state of strain a stress surface should exist if da is a complete differential.
If ^ = d<r/dt and .p = da/di this is the case when

If -If'
4* and <p, however, may also depend explicitly on the time t as in processes in which
softening or age hardening might be involved (i.e., some parameters entering in the
expressions for <pand might themselves depend on the time /) in the grain structure
of a metal while it is slowly deformed under a load. Comparatively little is known
quantitatively about <p and <pand how these quantities depend on o,e,i,t and on the
temperature.
Generally a theory of a plastico-viscous substance could be based on the equation

- /(yo,To) (28-31)

denoting with the subscripts zero the octahedral variables, thus postulating a general
stress surface in the sense which was just mentioned and assuming that -y0 and 70
express permanent shears. An equation like (28-31) presupposes that the flow stress
to is not dependent on the mean stress (<ri + <r2 + <r3)/3. It is furthermore under
stood that under this type of distortion no volume changes occur so that

«i + «2 + «o = 0 .»

1 One could alternatively consider cases in which either elastic or permanent or


both kinds of changes of volume of some form accompany the distortions in shear.
The case of a purely elastic dilatation in volume is treated in Sec. 28-7. In this con
nection a short note may be of interest published by A. It. Lee, M. Reiner, and P. J.
Rigden (Nature, Nov. 16, 1946, p. 706) quoting observations on the flow of concrete
in compression in which a permanent decrease in volume was recorded. A contrac
tion in volume of cast concrete during its period of curing is well known. It has been
investigated from theoretical angles by R. L'Hermite (Compt. rend, des laboratoires
du bailment et des travaux publics, Paris, 1945-1946, p. 1). It might be caused accord
ing to him through an increase of suction due to the capillary action of the water
exerted on the skeleton of rigid particles (sand) when the water gradually evaporates.
Permanent changes in volume might be caused in metal through aging independently
of the presence of stresses. Reversible changes might be due to temperature strains.
Thus there are a variety of physical causes which may produce permanent or reversible
changes of the volume. K. H. Swainoer (Nature, vol. 158, p. 165, 1946) found in
duralumin an increase of volume after cold work in tension. This is again made prob
able through the general mechanism of plastic distortion in metal crystallites after
severe cold work due to the formation of minute gaps and voids between the mozaic
blocks that have slipped within the grains, as G. Tammann pointed out long ago.
Measurements of density after severe cold work in metals even after compression
stresses have acted usually disclose a small decrease (a permanent small increase in
volume). Thus a number of phenomena are known which would not be covered
through Eq. (28-31). A step in the direction of further generalization of Eq. (28-31)
was recently suggested by A. Gleyzal (Dual Stress-strain Laws of Elasticity and
Plasticity,J. Applied Mechanics, June meeting in Buffalo, 1946) who proposes that
to = /(to,«o) and a„ = g(yn,m) be considered, the subscripts zero referring to the
octahedral planes of stress and strain.
420 THEORY OF FLOW AND FRACTURE OF SOLIDS

An interesting special case of the general law of deformation expressed by Eq.


(28-31) was proposed longer ago by Bingham in his work on the slow flow of certain
fluids1 (paints, suspensions). He characterized their behavior by postulating a
"yield value" (to*) for them, and he assumed furthermore that these substances dur
ing their motion behaved like perfectly viscous fluids provided that the flow stress
to > to*. These conditions are expressed by

to = ±to* + ny„ , to* = -\/(<ri* -a,*)« + ■ • ■,


7o £ 0 . (28-32)

The theory of this substance having a limit of plasticity should coincide in essential
respects with the theory of a perfectly viscous substance in those portions of a body in
which movements occurred. The yield value t0* and the viscosity coefficient m in
Eq. (28-32) are material constants. {<ri*,<rt*,<rt* are the principal stresses at which
flow starts.)
Ilyushin* and Ishlinsky3 applied the theory of flow postulated by Eq. (28-32) to a
remarkable case of an instability of a plastico-viscous equilibrium in a tensile bar.
Assuming that a bar of such a material has a profile containing regular shallow waves,
both authors investigated the condition under which the bar must start to neck locally
in the valleys of the waves. If the condition of instability is not satisfied, the waves
will be smoothed out. This work is the first attempt in which conditions were speci
fied under which a tensile bar of a material having well-defined properties must start
to neck down locally. Both authors developed this theory for necking for the plane
and for the rotationally symmetric problem.*

-+■
From physical considerations it must be said that the linear function r — t* py,
of the rate of shear which was assumed by Bingham and Ilyushin, probably over
y

simplifies the experimental facts even for those softer materials for which was

it
originally proposed and for which seemed to agree with the flow tests. It has been
it

noted that the true form of the creep function t = f(y) for solids at very small creep
rates requires that the curve representing should pass through the origin of the r,y
it

system of coordinates. Furthermore, for physical reasons t = f(y) should be an odd


function of and should have a steep but finite slope at the origin corresponding to
y

a
finite viscosity value dr/dy at very small rates of flow. The apparent knee the

in
t f(y) curve caused by
= its exponential character at larger shear rates must have led

Bingham, Eugene C, "Plasticity and Fluidity," McGraw-Hill Book Company,


1

Inc., New York, 1922. This discussed in detail in the tenth lecture, p. 133, M.
is

in

Reiner's book (loc. cit.).


Ilyushin, A. A., Deformation of Visco-plastic Bodies, Scientific Reports the
of
*

Moscow University, issue 39, Applied Mechanics, 1940. (Russian, English transla
tion prepared by Westinghouse Research Laboratories jointly with the David Taylor
Model Basin, Navy Department, 1947).
Ishlinsky, A. I., J. Applied Math. Mech. Academy Sciences U.S.S.R., Moscow,
3

of

vol. pp. 110, 405, 1943 (translated by Brown University and David Taylor Model
7,

Basin, 1946).
The principal axial stresses in the region midway between the valleys and crests
4

of the waves are slightly inclined with respect to the bar axis. This tends to retard
flow in the valleys while the mean axial stress across the minimum sections the
of

valleys, being larger than across the crests, tends to accelerate the flow there. When
the latter effect predominates, necking must continue.
SIMPLE AND COMPOSITE SOLIDS 421

observers to assume that this sudden knee in the curve represented a true "yield
value."1

28-6. The Elastico-viscous and the Firmo-viscous Solid. In numer


ous practical applications of the theory of elasticity, plasticity, or vis
cosity of solids it is entirely sufficient to introduce only one type of
strain, either elastic or permanent. The corresponding simple solids
were discussed in the previous sections and in Chaps. 25 to 27. In con
trast to these cases there are many problems of flow which require that
both the elastic and the permanent portions of the strains be simultane
ously considered. To dispose of a short term for designating substances
in which elastic as well as permanent strains must be distinguished such
materials may be called composite solids.2
To this second class of solid substances belong the elastico-viscous solid.
It was proposed by J. C. Maxwell in his classical paper, "On the Dynam
ical Theory of Gases" as follows:8 "A distortion or strain of some kind
which we may call t is produced in the body by displacement. A state
of stress or elastic force, which we may call <r, is thus excited. The rela
tion between the stress and the strain may be written <r = Et, where E
is the coefficient of elasticity for that particular kind of strain. In a solid
body free from viscosity a will remain = Et and da/dt = E dt/dt. If,
however, the body is viscous, a will not remain constant, but will tend to
disappear at a rate depending on the value of a, and on the nature of the

1 Similar remarks were made by M. Reiner on p. 137 of his monograph (loc. cit.)
in reference to the flow of the fluids which were the objects of the researches of Bing
ham and his pupils. The author, however, does not follow Reiner in his suggestion
that these fluids of a generally viscous nature be called "non-Newtonian fluids" as
distinguished from the viscous (he calls it a "Newtonian" or simple) fluid. This and
related terminology introduced by authors among the rheologists is awkward and
entirely unnecessary. (According to this recently suggested nomenclature, for exam
ple, the solid metals in their creep range should belong to this class of fluids. The
association of the names of certain investigators with simple mechanical properties of
materials is another practice which should not be encouraged, if for no other reason
than because the authorship of thought so frequently is doubtful in its origin. This
should prove less convincing if the name used is quoted in the negative sense, as in
the above example.)
* The term is somewhat ambiguous in the sense that it may be referred also to con

glomerates of two or more than two solid compounds joined in one solid having grains
of different chemical composition and mechanical properties of which trivial examples
are found among the alloys of metals or the natural rocks. It is not the purpose of
this or the following sections to analyze from a mechanical point of view the conglom
erates of elastic and viscous elements in a structure.
'Phil. Mag., vol. 35, pp. 129-145, 185-217, 1868 (quoted from p. 133).
422 THEORY OF FLOW AND FRACTURE OF SOLIDS

body. If we suppose this rate proportional to a, the equation may be


written
_~~ „ dt
f
~ a
da
dt dt

which will indicate the actual phenomena in an empirical manner. For


if t be constant
a = tne-"7-

showing that a gradually disappears."1


We may characterize an elastico-viscous solid also by stating that it
has a total strain t which is the
sum of an elastic («') and a perma
nent strain of a viscous nature (t")

<• = e' + t' (28-33)

and furthermore by assuming with


Maxwell that the elastic strain t
and the rate with which the per
manent strain changes, namely,
i" = dt" /dt are proportional to the
stress, e.g., in simple tension
t' = a/E and i" = a/3n, where E
is the modulus of elasticity
and (i the coefficient of viscosity
so that
dt _d£ W 1 da a
dt dt dt Edt +^l
(28-34)

Fio. 28-11. James Clark Maxwell (1831- After making use of the synthe
1879). {From "Great Men of Science," by
sis of small strains given in Chap.
P. Lenard, The Macmillan Company, New
York, 1933.) 24 and under reference to Eqs.
(24-7) expressing Hooke's general
ized law for an elastic and Eqs. (26-7) for a viscous material, an elastico-
viscous solid is generally characterized by the equations:

D-
Ut = 5f T
+ £ t =
+ av
ZK
+ ff,
(28-35)
2G 2n

of the rate of stress, and of


D„ D„ and D; are the deviators of stress,
the resultant rate of strain, G the modulus of rigidity, and K the bulk
trans
' During this of unloading an originally elastic strain is gradually
process
formed in a permanent strain. This phenomenon is called the relaxation of stress.
SIMPLE AND COMPOSITE SOLIDS 423

modulus. After combining Eqs. (28-35) with the conditions of equi


librium [Eq. (25-4)] a system of linear differential equations is obtained.
As a second example of a composite solid the firmo-viscous solid pro
posed by H. Jeffreys1 is quoted. In contrast to the former case the stress
a in simple tension is now assumed to have additively an elastic and a
viscous part a' = Ee and a" = 3jt« so that

ff = „• + a" = Et + 3m jt •
(28-36)

Let I be the length, A the area of cross section of an elastic bar held in
vertical position fixed at one end and connected with a heavy mass m
at the other end and let x be the small displacement of the mass, measured
from its static equilibrium position when it is excited to small oscillations.
An elongation x causes an elastic reaction Ao = AEx/l. If internal
damping of a viscous nature is also present the reaction will be

p = Aa _ A[Ex + 3/i {dx/dt)\

and the equation of motion of the mass

-~p„ dx
£ + (£)r + (45)~°-
d2x
"' <**>
mw< X
This is the well-known equation of the damped oscillations of a mass
attached to an elastic member when a damping force like the one produced
by a dashpot is present. Thus we note that the "firmo-viscous" sub
stance defines an elastic medium in which considerable internal damping
forces of a viscous nature are excited under rapid oscillations. The
inertia terms in the dynamic equations of motion of such a substance must
be considered.
H. Jeffreys in Cambridge, B. Gutenberg2 in Pasadena, and others have
utilized the theory of these two solids for the mathematical treatment of
a number of most interesting geological phenomena in the upper and in
the deeper strata of the earth crust [postglacial uplift of the region of the
Great Lakes in the United States and Canada, the problem of the folding
of mountain chains (Jeffreys), the formation of the rock-salt domes in
Louisiana (see papers by American petroleum geophysicists), seismic
phenomena (Gutenberg, etc.)]. Jeffreys points out that the "relaxation

1 "The Earth," 2d ed., 1929, Cambridge. Also Proc. Roy. Soc. London, ser. A,
vol. 138, p. 283, 1932, and other papers.
» "Physics of the Earth," vol. VII, Internal Constitution of the Earth, McGraw-Hill
Book Company, Inc., New York, 1939, and other papers.


424 THEORY OF FLOW AND FRACTURE OF SOLIDS

time" (T = 3n/E) for the elastico-viscous solid is a large and for the
firmo-viscous solid a small quantity. The viscous term in the latter case
becomes significant at extremely large rates of strain (under the oscilla
tions of a violent earthquake), in the former case only after elapse of long
(geologic) times.

28-7. Plastic Deformation Considering the Elasticity of the Material. On loading


of material bodies having a sharply defined yield point and a distinct elastic range a
little beyond the point at which the plastic limit in certain portions of the bodies has
been reached in these regions, the permanent strains t",y" will be of the same order
of magnitude as the elastic strains i',y'. In such cases it will be necessary to consider
the elastic deformation as well as the permanent one. Under the assumption that no
finite rotation in the principal directions of stress accompanies this deformation and
that both strains are small, the expressions for the resultant strains t,y consisting of
the sums of their elastic and permanent parts were developed in Eqs. (24-11) and
(24-12):
t* = 4>Wz - v(<ry + <r,)\ , • • • y„. = 2(1 + r)+ry. , ■■■
(28-38)

<f>
In these six relations two unknown functions appear: the flow function and the
variable Poisson's ratio which was denned by
v

In addition to the equations listed in Chap. 27 denning the distortion of an ideally


plastic substance with purely plastic strains one more equation now needed the

if
is
corresponding equations for the resultant strains are now considered together with
the conditions of equilibrium and of plasticity. This furnished by the condition
is

that »" = J^, i.e., that the permanent portions of the normal strains do not contribute
to the volume dilatation in the plastic zone1 or that the total dilatation of volume there
remains elastic and equal to

«* + «v + u = 3«r (28-40)

where a denotes the mean stress (<r, + «•» + a,)/3 and = 1/3K, K being the bulk
k

modulus of elasticity of the material.


An equivalent equation expressing that the total dilatation of volume elastic and
is

proportional to the mean stress a Eq. (24-16)


is

- 2v)* - (28-41)
(1

«
<j>

connecting the flow function with the variable unknown Poisson's ratio and furnish
v

ing the last missing equation.


It was pointed out in Chap. 24 that Ilyushin' deserves credit for having eliminated
through introducing a variable v ratio certain difficulties which had been noticed in

This justone plausible assumption, derived from experiments on which to hase


1

is

elastic-plastic stress-strain relations. Eor other possible assumptions, see footnote


on p. 419.
Loc. cit.
8
SIMPLE AND COMPOSITE SOLIDS 425

problems of the partial yielding of material bodies.1 These stress-strain relations


Eq. (28-38) are expressed in the deviatoral condensed form for the resultant strains by

D, - (1 + v)<t>D, (28-42)

corresponding to Eqs. (24-18) and (24-21).


If the strains «" in the yielded zone have become predominant, in these regions v
approaches the value \^. The "elastic" zone must be surrounded by a layer of
material in which the adjustment of the values of Poisson's ratio takes place from the
value v = H (for steel) corresponding to purely elastic strains to the value v = %.
Although the preceding remarks may apply to simpler cases of partial yielding such
as the ones encountered in bending of beams or in other cases it should be noted again
that elastic-plastic stress-strain relations in finite form like Eqs. (28-38) must be
replaced through the corresponding relations which are valid for the infinitesimal
increments of strain when the components of stress under which elements of material
yield change their values during the plastic distortions. This occurs when a plastic
zone advances through a body carrying with it its own field of stress (even although
in certain simpler applications the principal directions of stress do not turn in the
elements of material). In such problems the increments of total strain must be con
sidered, which are equal to the sums of the increments of the elastic and of the perma
nent portions of the strains necessitating a step-by-step integration of all stress-strain
relations in addition to the integration of the other equations. How this may be car
ried out has been described in a paper by R. Hill, E. H. Lee, and S. J. Tupper,1 and
the integration of the infinitesimal increments of total strain has been extended by
K. H. Swainger to the case of a strain hardening metal. He dealt in one investigation
with strains of small' and in another investigation of finite* magnitude, suggesting
also that the computations for three-dimensional states of inhomogeneous stress
might be simplified by replacing the stress-strain curve by two straight lines, one of
which describes the elastic and the second, having a much smaller slope, the plastic
range of strains.6 Special reference to the ideas proposed for dealing with problems
involving incremental stress-strain relations in these afore-mentioned papers is made.

K. H. Swainger, London, in a letter written to the author in 1943 in reference to


1

the problem of partial yielding in metal cylinders, pointed to the difficulty arising
along an elastic-plastic boundary if Poisson's ratio is assumed to change discontinuously
from the value 3^ to J^. The discontinuity of certain stress components resulting
from the former assumption is eliminated if Eqs. (28-38) to (28-41) are used.
J The Theory of Combined Plastic and Elastic Deformation with Particular Refer
ence to a Thick Tube under Internal Pressure, Proc. Roy. Soc. London, ser. A, vol. 191,
No. 1026, Nov. 18, 1947.
* The Compatibility of Stress and Strain in Yielded Metals, Phil. Mag., ser. F,
vol. 36, pp. 443-473, 1945.
4 Stress-strain Compatibility in Greatly Deformed Engineering Metals, Phil. Mag.,

vol. 38, pp. 422-439, 1947. See also "Severe Deformation," 7th International Con
gress of Applied Mechanics, London, September, 1948.
s The latter author introduces a new definition for finite normal strain, assuming it

equal to the difference of the stretched and initial length divided by the stretched
length. This new normal strain is made equal to (1 + e — 1) : (1 + «) = «:(1 + e)
(e denotes the conventional normal strain).
CHAPTER 29

CONSTRAINED FLOW OF IDEALLY PLASTIC MATERIAL.


STRESS TENSOR TRAILS RESULTANT STRAIN TENSOR

29-1. Constrained Flow Neglecting Elasticity. In the preceding cases


of flow finite relations connecting the components of stress with those of
the tensors of strain or rate of strain could be postulated because the
assumption was made that the principal axes of stress would not rotate
relative to the elements of the material when the strains increased. In
the following pages we shall devote some attention to the integration of
infinitesimal increments of elastic and of plastic strain while the tensor of
stress remains continuously at the limit of plasticity but changes the
directions of its principal axes in the elements of material. This must
occur if certain kinematic conditions are imposed on a body yielding under
loads through solid connections with other bodies, which do not permit
it to yield in that manner in which it would deform under the same
system of stresses if the boundaries were unrestricted in their motion,
e.g., if the resultant strains were prescribed along the boundary of the
body and restricted in a given manner in their free development.
Suppose that a block of plastic material is deformed under a homo
geneous pure shear az = —<rv, a, = 0; tw; rx, = t„x = 0. Consider the
increment of this pure shear:
<f>

dtz = = — dty , dy^ = 3d t„ (29-1)


~2~ax
,

<f>

where all the strains are permanent, dtt = dyzl = dyyt = and denotes
0

the flow function.


On the differentials dtz and dy^ a condition may be imposed. For
example, we may prescribe that

<x
=
/(X) Txv
= g(\) (29-2)
,

(which could be the time


a f),

where" variable parameter so that


is
X

dtz = f'(\) d\, d\. These equations prescribe


dy^y = g'{\) certain
deformation which may be represented by curve tz = Fiy^) an
in
is a

«i,7i» plane of coordinates [this last equation obtained by eliminating


0, X

from Eq. (29-2)]. This curve in general must begin at the origin (tz =
7i» = the material deformed from an unstrained condition.
is
0)
if

426
CONSTRAINED FLOW 427

If this
curve is a straight line originating at the point t, = 7„ = 0
we obtain the case of unrestricted flow without a rotation of the principal
axes of stress, since when t,/y,y = c = const, dt,/dy,y = o,/^,y = c and
from the condition of plasticity1

+ =
<r,2
^ (29-3)

(<X0 yield stress in tension) on account of the constant ratio a,/2T,y it


follows that <r, = —<fy as well as Txy are constant values during the flow
and after integration of Eqs. (29-1)

u-^Zl, «»=-^r' (29-4)

the stress-strainrelations for a pure shear are obtained.


If, however, t,
= F(y,y) is not a straight line passing through the origin

we are led to an example of constrained flow. Substituting the values of

taken from Eqs. (29. 1) in F,q. (29-3) we obtain for the increment d<p:

d* = —\— - V4(rf02 + (dy,y)2 (29-6)

and for the flow function:

and the stresses:

*' _ t" jro


2<ro 1 1
— - (29-81;
VsV^fW,JdJ,)2' V3 VWtJdy,y)* + 1

Since the ratio a,/r„, = "2dt,/dyxy is not constant the principal axes of
stress must continuously change their directions in the body during
this deformation.
The cases of unrestricted and of constrained flow may be visualized
if a point P(<ri,ai,a3) on the general surface of yielding

(<ri
~
<r*Y + (<r2
- <r3Y + (<r3
- <ri)2
= W (29-9)

is considered representing the state of stress in a certain material element.

1
Equation (29-3) is obtained from the general condition of flow, Eq. (27-9) after
substituting in it Tvi = ti, = a, = 0, <fy = —a,.
428 THEORY OF FLOW AND FRACTURE OF SOLIDS

Equation (29-9) represents a circular cylinder in the system ai,at,a3 of


rectangular coordinates.
If point P remains in a fixed position on the cylinder the element can
deform freely. On the contrary, if the flow is constrained, the point P
will be forced to change its positions continuously on the surface of the
cylinder.1
29-2. Constrained Flow Considering Elasticity. Let us now describe
the cases of constrained flow by including also the elastic strains in the
consideration. In accordance with previous practice (Chap. 24) denote
the elastic strains by one, the permanent strains by two primes and the
resultant strains, which are the sums of the former ones, without a prime.
The components of the elastic strain satisfy Eqs. (25-2). Suppose that
the components of stress have reached values under which the material
yields. These stresses must then satisfy the condition of flow, Eq.
(27-9):

(<r. + a,)2 + (er, - <r,)2 + (<r*


- axY + 6(rw» + rVI2 + r„)J
= W. (29-10)

Consider an infinitesimal increment of the resultant strain. When this


is produced the stresses may change by the increments d<rx, dav, da,, drz„
drvz, dr,x. These six quantities must satisfy the equation which is
obtained by differentiating Eq. (29-10) or the equation:

— — —
[vu

do,
!L"2

day
*—^
+

( or
L"2

')

dax
f

(<r.
"J
)

+ 3(t„ rv, dry, r„ dr„) =


+

dTxv 0 . (29-11)

The elastic strains change by the increments

dt,' = - • • ■
+

4>'[d<rx v'(doy da,)],


,

dyv/ = 2(1 • ■
+

v')4>'dTy„

(29-12)

One may think of representing also the states of plastic strain by a point having
1

the principal strains «i,«i,«! as W. M. Baldwin suggested (Metals Tecknol., 1946)


in
a

system of rectangular coordinates ei,ej,ci. For incompressible materials the condition


«i + ci + «a = represented by plane passing through the origin of the coordi
is
0

nates which perpendicular to one of the space diagonals. If both systems


is

«i,e2,«3
and («i,«,«») are made to coincide with their axes, this plane perpendicular
is

(<ri,(Ti,irt)
to the cylinder representing the surface of yielding Eq. (29-9). If point
P

(ffi,ff:,oj)
situated on the cylinder projected on this plane point P0 obtained in which the
+ is
is

generatrix passing through intersects the plane «i + a t% = 0. The straight


P

line connecting the origin with P0 (situated in the plane <i + <i «» "•
+

represents
0)

the states of plastic strain for unrestricted flow, as Baldwin shows. Any curve drawn
on this plane would define succession of plastic states under constraint.
a
CONSTRAINED FLOW 429

Assuming v" = the plastic strains simultaneously increase by the


}-£
amounts

<U." = d<t>"
L, - ^—^Y • • •
, dyv." = Sd<p" tv (29-13)

These last equations conform with Eqs. (27-1) and with the requirement
that the deviator of these increments of the permanent strains is coaxial
with and its components are proportional to those of the deviator of
stress.
If the increments of the resultant strains

dtx = dt/ + dt," ,


■ ■ ■
, dyy, = dyj + dyv," ,
• • •
(29-14)

are known, a sufficient number of equations is available together with the


condition (29-11) from which both the dtj, dey, . . . , dyyi', . . . and
the dtx", d(v", . . . , dyyt", . . . may be determined. We note, how
ever, that the deviator of the increments of the resultant strains will not
be more coaxial in general with the deviator of stress.
L. Prandtl1 and A. Reuss2 first called attention to such constrained types
of flow. The former dealt with the case of plane strain showing that in
this particular case the addition of the two tensors of the elastic and
plastic strains may be reduced to the problem of a geometric addition of
certain coplanar vectors. The latter developed the general equations
for this case, expressing them, however, not for the increments of the
strains, but for the components of the rates of strains. This made it
necessary for Reuss to integrate his equations first with respect to the
time t. The equations which were given above in the text avoid this
integration. They are essentially those by A. Reuss.
29-3. Example of Case of Combined Tension and Torsion. Let us
consider for an example the different modes of flow in which a material
element may respond under a simple system of stresses. We choose for
this purpose a tension-torsion test, a case to which Reuss also inde
pendently later applied his equations.3 Assume a brick-shaped element
of material which is stressed under the combined action of a tensile stress
a and a system of shearing stresses t. The faces of the element are
parallel to the xy, yz, and zx planes. A normal stress ay = <j in the y
direction and the shearing stresses tvi = t^ = r act on the element. All
1 Proc. Intern. Cong. A pplied Mechanics, Delft, 1 924.
» Z. angew. Math. Mechanik, vol. 10, p. 266, 1930.
3 Around 1926 a tension-torsion machine of a special design was constructed at the

suggestion of the author by A. Sonntag and built by the Ixjsenhausen Works in Dtis-
seldorf . It was contemplated to use it under different loading conditions for observa
tions of noncoaxial elastic and permanent strains (see Fig. 17-12, p. 247).
430 THEORY OF FLOW AND FRACTURE OF SOLIDS

other components of stress are zero. The wall of a steel cylinder under a
tensile force acting in the axial direction and under a simultaneously
applied torsion moment around the axis would be stressed in this manner.
a. Flow When Components of Stress Tensor Remain Constant. Suppose
that we neglect first the elastic components of strain. The state of
plastic strain in the element is expressed by

Cy
= t , €x
= «z
= — > y«z = y , 7*V
= Yri = 0 . (29-15)
g

The plastic stress-strain equations according to Eqs. (27-1) in Chap. 27,


assuming that Poisson's ratio v" = J£, are expressed by

t = fa , y = 3<£r . (29-16)

They contain implicitly the condition that the tensors of stress and of
permanent strain are coaxial (and that the deviators are proportional).
After dividing the second by the first equation we note that coaxiality is
preserved, provided that

*-£. (29-17)*

At the limit of plasticity a and r must satisfy Eq. (29-10) which becomes
„■>■ + 3T2 = aJ = const . (29-18)

In a and r this is the equation of an ellipse.


the coordinates A state of
stress (o,t) at this limit is represented by a point P of this ellipse (Fig.
29-1). The states of plastic strain may similarly be represented by a
point Q(t,y) in an e,y plane. While P remains in a fixed position on the
ellipse, point Q according to Eq. (29-17) must describe a straight line
starting at the origin 0 in the e,y plane. The slope of this line y/t must
be equal to 3r/<r or equal to three times the slope r/a of the line OP in
Fig. 29-1.
b. Same Case Including Elastic Strains. Suppose now that we wish
also to include the elastic portions of the strains. As long as point P
remains in the interior of the ellipse [Eq. (29-18)] the strains are elastic
and equal to

*' =
r y, =
h' (2M9)

*
Condition (29-17) is obtained also by computing the angles a and 0 under which
the principal directions of stress and of strain are inclined with respect to the axis I
from [see Eq. (10-4)]
2t -v
tan 2a = i tan 20 = —
*z <»
after making j3 = a.
CONSTRAINED FLOW 431

Let e0 = <ro/E be the elastic strain in tension when the yield stress ao
is just reached. When a pair of values a,r reach the limit of plasticity
[Eq. (29-18)], the elastic strains t' and 7' must satisfy the equation
3
+ (29-20)
4(1 +
This is again the equation of an ellipse in the t',y' plane [e0 and

2(1 + v')*o
V3
are its semiaxes]. The states of per
manent strain are represented by
points Q outside of this second el
lipse. In Fig. 29-1 an additional
ellipse is shown by a heavy line
which would be obtained if V = %
(it has \/3 «o for its vertical semi-
axis) . We note that if v' is not equal
to J^, but say equal to as for steel,
the two ellipses do not differ too much
from each other. In the following we
shall use only the ellipse correspond
ing to the value /= J^, namely,

+ = eo (29-21)

Thus when <t,t remain at a level at


which Eq. (29-18) is satisfied point Fio. 29-1. Representation of path of
moves on a straight line outside straining for case of combined tension and
Q torsion.
of the ellipse [Eq. (29-21)] on the
extension of the radius vector OQQ of this ellipse. If the flow is inter
rupted, e.g., in Qi> and the stresses reduced to zero, point Q moves back
as the material contracts elastically on the same line to point O\, until
O\Qi = OoQo- The components of the elastic, plastic, and resultant
strain appear readily in the coordinates of points Qi and Oi and in their
differences (Fig. 29-1).
c. Let us finally consider a case in which the principal axes of the tensor

of stress rotate. Instead of maintaining a constant value of a and t we


assume on the contrary now that after these stresses have reached the
limit of plasticity with the values <r1 and n point Q is moved along some
given curve in the t,y plane, e.g., along the straight line Q0Q1 shown in
Fig. 29-2. This would occur in a tension-torsion test with a hollow
432 THEORY OF FLOW AND FRACTURE OF SOLIDS

cylinder if subsequently the specimen were stretched at some constant


rate and simultaneously twisted at a given rate. We wish to find the
loci of the points O\ (t",y").
Suppose that the total strain has increased by the increments dt and
dy corresponding to the infinitesimal vector Q0Q\ in the e,y plane (Fig.
29-3). Resolve the vector into its components Q0R and RQi parallel to
the directions of the radius vector 00Qo and of the tangent line to the
ellipse in Q0. These two vectors represent the increments of the plastic

Fio. 29-2. A case of constrained Fio. 20-3. Increments of the elastic


flow represented by the elastic, and plastic strains,
plastic, and resultant strains (case
of combined tension and torsion).

and of the elastic strain. The first conforms with the rules of flow which
were just explained for an unrestricted permanent deformation. The
second (elastic) component can be produced by displacing point Q along
the element of arc of the ellipse moves on the "inside" of the ellipse
(if
Q

[Eq. (29-21)] very near to this arc this constitutes only elastic changes
in

strain). Any subsequent increment of the total strain (dt,dy) can thus
be resolved into its two components in an analogous manner as a force or
a

velocity vector resolved into two components along known directions.


is

The condition that point moves on any given straight line


is

expressed
Q

by
= = const
(29-22)
c

at
where
dt = dt' + dt" = da a
+

d<t>"
,

(29-23)
dy = dy' + dy" = dr +
r

3(</>' d<t>")
.
CONSTRAINED FLOW 433

Furthermore, from Eq. (29-18),

a da + 3t dr = 0 . (29-24)

As just stated, the example chosen is an illustration of a special case of


the superposition of two strain tensors in space which follows the rules of
the geometric addition of two vectors in a plane. The coordinates of Q
are the resultant strains t,y; those of 0 are the permanent strains (e",y");
their difference represents the elastic strains (e',7')- When Q moves along
a given path in the e,y plane, point 0 trails its movement in a manner
reminiscent of some problems describing curves of pursuit in elementary
differential geometry.1 In fact, move a pin representing point Q along
its prescribed path, this pin fitting in a slot which has been cut in a solid
plate (covering the plane) along the ellipse A0QoB0. The pin will dis
place the plate when it moves. If the plate is prevented from making
rotational movements, the center of the ellipse describes the locus of
the points 0. It trails Q at a distance equal to the variable length of
the radius vector OqQo (because the triangles QoOoS and ROiQi are
congruent).2
Thus we see that the plastic and the elastic strain tensors (the latter
is coaxial with the stress tensor) figuratively speaking must in general
trail the resultant strain in such a manner that they all tend to align
their principal directions, parallel to a common direction. When the
path of point Q is a straight line such as QoQi in Fig. 29-2, these two ten
sors must soon align themselves into positions in which all three become
coaxial tensors. This occurs within total strains which within practical
accuracy are three to four times larger than their elastic portions, as can
be judged from the constructed curve of the points 0 in Fig. 29-3.

The four equations (29-22) to (29-24) define a differential equation for the unknown
flow function <j>":
3r[*'(di-/r) + d*")
(29-25)
6[0'(rWa) + (!<(,"]

If c = 0 or for a distortion occurring under a constant total unit shear, y = const,


evidently

- -
<*>'

«." In (29-26)

and = « or when = const,


if

e
c

<t>" - 0'ln--
<j (29-27)

Example: a dog pursuing a rabbit.


1

Through a change of coordinates using an affine transformation one may deal with
1

circle instead of the ellipse, as was suggested for the plane problen by L. Prandtl.
a
434 THEORY OF FLOW AND FRACTURE OF SOLIDS

Equation (29-25) can be integrated after introducing an angle a as a new inde


pendent variable defined by

a = <ro cos a , t = —5= sin a (29-28)


V3
which expressions satisfy Eq. (29-24). (a is the angle TOC in Fig. 29-1.) Equation
(29-25) is then transformed into

— = — cotan (a —
a0) (29-29)
da

provided that the constant a0 is taken so that

tan a0 = -4= (29-30)


V3
and that ao thus measures the slope of the line QoQi in Fig. 29-2. The integral of
Eq. (29-29) is

= In
♦»
sin (a — ^
ao)
, (29-31)

where ai is the angle cr for which Eq. (29-28) furnishes the initial values of the stresses
<r = <ri and t — n. W ith these expressions the total strains are determined and found
equal to
« = «o[cos ori + /(«) cos or0] , 1
(29-32)
7 = V3 «o[sin ai + /(a) sin a0] , J

where /(a) denotes the function

f{a) = In tan ,-. — T-jzi (29-33


[(a a°) /2\

and «o = <ro/E = <t>'<r„. The expressions for the elastic strains are determined by using
Eqs. (29-28) and are found equal to

«' " = «o COS a , )


y> = 3^'T = V
3 i. sin « , J
1-W-34J

and the permanent strains t",y" can be computed from the differences t — «' and
— y'.
7
For the special case of a distortion occurring under a constant unit shear (->• — const)
our formulas give

f . . tin (a,/2)"| /nn


•-'•L^a.+ln t;ur(a/2) J (29-35)

and if one assumes the initial stresses ai = 0, n = <ro/\/3 the curve representing the
increase of the strain « with the stress <r is expressed by the function

+
t = ^° In ) —
& I ^."l
(a /ao)
= eo tanlr> (<r/<r„) (29-36)

in agreement with the theory by Reuss.


CONSTRAINED FLOW 435

Applying these results to a hollow cylinder of steel that has been brought under the
mutual action of a tensile stress ai and a shearing stress n under which it would just
start to yield, we see that if the cylinder is prevented from any further twisting, e.g.,
by holding y = const, the shearing stress t must start to drop and simultaneously
the tensile stress a to increase until finally r = 0 and a becomes equal to o-0. The
initially elastic unit shear y' = 3^>'ti is gradually transformed into a permanent shear.
Finally the cylinder will stretch only by pure tension and will preserve its initial
twist even after the test is interrupted and the tensile load removed.
Some experiments of this kind with steel tubes have been carried out by K. Hohe-
nemser1 using the tension-torsion machine mentioned above and further discussed by
him and W. Pragcr.* Their results, however, were not too convincing. (They seem
ingly supported the theory of unrestricted flow.) Combined tension-torsion tests with
hollow cylinders, in which other questions were investigated, however, were made on
flow by G. I. Taylor and H. Quinney3 and on fracture by E. A. Davis.4

1 Z.
angew. Math. Mechanik, vol. 11, p. 15, 1931.
»Z. angew. Math. Mechanik, vol. 12, p. 1, 1932.
1 Trans. Roy. Soc. London, ser. A-V, vol.
230, p. 323, 1931.
« Trans. ASME, 1940.
CHAPTER 30

THEORY OF THE PLASTIC DEFORMATION OF


THICK-WALLED CYLINDERS. CYLINDER OF
IDEALLY PLASTIC MATERIAL
In a hollow cylinder stressed symmetrically around its axis and uni
formly along the length, the radial, circumferential, and axial directions
are principal directions of stress and strain. It will be in order, as in the
mathematical theory of elasticity when the two-dimensional problems
are treated, to distinguish those rotationally symmetric states of plane plastic
strain in cylindrical bodies in which the axial strain is constant throughout
the cylinder as one group and the states of plane plastic stress for which the
normal stress in the direction parallel to the axis of the cylinder vanishes as a
second group. The first group embraces the distributions of stress and
the distortion in long cylinders, the second of flat circular disks or of
rings stressed parallel to their middle plane. Within each of the two
groups one is interested, in applications, in considering cases in which
the strains are either infinitesimal or of finite magnitude. Because of the
importance of the ductile metals and of their alloys as materials of con
struction, it will be in order to analyze the flow of a cylinder of an ideally
plastic substance (representing the case of a metal having a sharply
determined yield point) and of a metal deforming under monotonously
increasing yield stresses (of a strain hardening metal). A variety of such
cases are encountered in practical applications, when hollow cylinders
are permanently deformed through high internal or external pressures,
during the forming operations of tubes through rolling, or when tubes of
softer metals are extruded through slightly tapered cylindrical orifices,
etc.
It has been pointed out also that, in heavy-walled cylinders of those
metals which have a well-defined yield point at which the permanent
deformation suddenly begins, elastic and plastic regions must be distin
guished and that in such cases the volume elasticity should be considered
within the permanently distorted region. These cases of the partial
yielding in cylindrical bodies therefore require a synthesis of small elastic
and permanent strains of comparable magnitude and these cases will
occur during the gradual transition from the elastic to the plastic states
of strain. After the permanent strains have further increased, the bodies
436
THICK-WALLED PLASTIC CYLINDERS 437

are essentially in the completely yielded condition. Various cases of


this kind will be treated in this and the following chapters.1
30-1. State of Plane Strain in Ideally Plastic Cylinder. The assump
tion will first be made that the entire cylinder is in a plastic state of
equilibrium under the action of uniform pressures which may be exerted
on either the inner or the outer cylindrical surfaces and that in addition
normal stresses of the required amounts act in the axial direction and are
distributed so that they maintain the plastic state throughout the wall of
the cylinder.
Let the normal stress in the radial, tangential, and axial directions be
denoted by <rr,<Tt,<i,, the permanent conventional strain in the same direc
tions by er,«<,«j, and the natural strains by I, = In (1 + er), e< = In (1 + tt),
e, = In (1 + «.). With the exception of the axial strain t,, which is equal
to a constant value «0 over a section normal to the axis of the cylinder, the
strains tr,tt and the stresses will depend on one independent variable,
namely, the distance r' of a point from the axis measured in the deformed
condition of the cylinder. If strains of finite magnitude are considered,
it will be in order to distinguish the distance r of a point in the original
unstrained condition of the cylinder and the corresponding distance r'
of the same material element in the strained cylinder, similarly the inner
and outer radii a,b before and a',b' after the distortion has occurred.
Assuming an incompressible material, when the strains are infinitesimal,
we have
«, + u + u = 0 . (30-1)

If p denotes the radial displacement, since er


= dp/dr, et
= p/r,

£j = «o = const

after inserting these values in Eq. (30-1) and integrating, the displace
ment is found to be

1 The pressuresrequired for permanently deforming heavy-walled cylinders through


internal cold working have been computed in many papers since the times when the
Austrian general Uchatius discovered the process of strengthening the bronze barrels
of field artillery guns by internal cold working, a method which since then has been
broadly applied in the arsenals making gun barrels out of steel. Many of these invest i-
gations on flow in cylinders have been carried out by artillery officers in various
countries and cannot be quoted in detail here. Recent references may be found in
papers by N. M. BELAiEvand A. K. Sinitski (1938); W. W. Sokolovsky (1943); and
others, dealing with the problem of the plastic distortion of cylinders (published by
the Academy of Sciences of the U.S.S.R., Moscow, available in English translations
prepared by Brown University and the Westinghouse Research Laboratories); by
R. Hill, E. H. Lee, and S. J. Tupper (1947); and by C. W. MacGreoor (1948), to
be quoted later.
438 THEORY OF FLOW AND FRACTURE OF SOLIDS

(30-2)
p
=-<u|+£
where c is an integration constant.
The permanent infinitesimal strains er and «< (satisfying the condition
of compatibility) in the cylinder are therefore given by the formulas

«r - _ to
"2
£
r2
'

(30-3)
««
= —(0s +
, C
~i

Obviously, the character of the plastic distortion in cylinders will depend


for example, both t0 and are positive constants, for
if,

on the ratio toai/c;

c
large values of toa2/c the cylinder stretched predominantly by axial

is
stresses a, in tension, has small value by radial and tangential
if

«oa2/c
stresses, etc. a
If, on the other hand, the strains reach finite magnitudes, we shall make
use of the natural components of strain i„h,h which satisfy the condition
of incompressibility
+ =
+

(30-4)
h

0
h

portion of the cylinder between the radii and having height


A

a
r
equal to the unit of length, after distortion will become a cylinder having
the radii and and a height + e„ = e*°. Since the volume does not
a'

r'

change, we must have


4- «0)(r" - a") = e'\r" - a") = rl - a* (30-5)
(1

Considering r,a,b as fixed radii, Eq. (30-5) defines the positions of the
corresponding points r',a',b' in the deformed state, and we may consider
r',a',b' as functions of the time when the distortion increases. The
t

tangential and radial natural strain and are equal to


i,

lr

- - s In r"
r'

= In = In
+
c«)
(1
h

—t

= — «o — = — «o — „ In -j
h

.
h

Denoting by a dot differentiation with respect to the time the natural


/,

rates of strain which may be denoted by the letter u are


= Ut = u, = «, (30-7)
i,
ir

wr
.
,

After differentiating Eq. (30-5) with respect to the time we obtain for
t,

/f = a>a' - Uo{r"
~
"^
(30-8)
THICK-WALLED PLASTIC CYLINDERS 439

and may compute the rates of strain from Eqs. (30-6) using (30-5) and
(30-8):

Ul ~
f" _ Uv
+
2a'd'
' + m0«"
7 2" 2~77T_ (30-9)

ur = — u0

^
ut .

After denoting by Co the quantity not dependent on r' :

c„ = (30-10)

it is noted that the natural rates of flow ur and u, during the process of a
finite radially symmetric distortion of a cylinder

">■
- Mo

2
Co

r" '

ui- - Mo
g-
,
+ y.
Co
,
(30-11)

are expressed as functions of the radius r' in a form entirely similar to the way
the infinitesimal strains «r and tt were expressed in Eqs. (30-3) as functions
of the radius r.
As long as the strains remain infinitesimal, it will be sufficient to intro
duce the plastic stress-strain relations by three equations of the form

(30-12)
ar=*(CTr-^±^)(etc.
after interchanging the subscripts r,t,z in cyclic order. When the strains
reach finite magnitudes these relations cannot be used, however, and for
them must be substituted three stress-rate-of-strain relations of the form

= ir = - ^p)
(ar

Mr * etc. (30-13)
,

in which <rT,<Tt,tr, are the components of stress acting on an element of


<f>

material in its deformed condition. The flow functions and for an


\f/

ideally plastic substance are pure functions of the state of strain (er,e,,€,)
and of the rates of strain {ur,ut,ut), respectively.
30-2. Ideally Plastic Cylinder without Axial Extension. When the
axial strain vanishes, e, = and uz = from Eqs. (30-12) or (30-13)
0

expressing either e, or u, follows that


it
440 THEORY OF FLOW AND FRACTURE OF SOLIDS

and from the condition of flow

(a, ~ arY + (<r,


- CtY + (*, - cr,)2 = W (30-15)

where <ro is the yield stress in simple tension, therefore that

<ft
- - ^ <rr ± = ±k. (30-16)
V3
After substituting this value of the stress difference in the condition
of the equilibrium of the stresses <rr and <r,

r' *L' = *t
- ar = ±Jb (30-17)
dr

the radial stress is found equal to

= ±k In r' + ci
<rr (30-18)

where In denotes the natural logarithm and Ci is an integration constant.1


The boundary condition for a hollow cylinder carrying an internal
pressure p on the surface r' = a' is expressed by r1 = 6', o> = 0 and
similarly, in the case of external pressure acting on the surface r' = b'
the boundary condition is r' = a', o> = 0. In the first case, since for
r' = upper and, in the second case, since for = a', <rt <
b',

ci > the
0,

r'

0
the lower sign must be chosen in Eq. (30-16). In the first case Ci =
— In b'; in the second Ci = In a'.
fc

The distribution of the stresses therefore given by the following two


is

groups of formulas:
Since <r,according to Eq. (30-14) the arithmetic mean of ov and <rt, the two smaller
*

is

Mohr's circles for the state of stress in any material element have equal diameters.
We note that in this case (e2 = 0), the state of stress in a cylinder, a pvre shear on
is

which pressure equal to <r,. The points representing the


is

superposed hydrostatic
a

plastic states of stress crr!a,,a, in hollow cylinder according to Eq. (30-16) are located
a

along two opposite generatrices along the circular cylinder representing the sxirface of
yielding f(aT,a,,a.) = in Eq. (30-15).
0

We note that in plastic cylinder only one boundary condition can be prescribed,
1

in contrast to the cases of elastic cylinders, for which two such conditions may inde
pendently be prescribed, e.g., one on the inner and the other on the outer surface.
This difference in the property of the solutions caused by the different character of
is

the differential equations expressing plastic and elastic equilibriums, to which refer
ence will be made later.
THICK-WALLED PLASTIC CYLINDERS 441

Cylinder under Cylinder under


Internal Pressure1 External Pressure

(30-20)

We obtain the pressure p under which the cylinder yields by computing


the value of the radial stress a, for
r' = a' or r' =
b',

respectively, and
find that in both cases the pressure p

r
equal to
is

2*
%.

= ln = In (30-21)
fc

%
p

The stresses are represented by


three equidistant logarithmic curves
(Figs. 30-1 and 30-2). The circum
ferential stress at reaches its largest
(absolute) value at the outer surface
= at = for the cylinder under
b',
r'

internal pressure and at the inner sur


face = a', at = — for the cylinder
r'

under external pressure.


EXTERNAL INTERNAL
The resultant force transmitted by PRESSURE PRESSURE

the wall in axial direction Fio. 30-1. Fio. 30-2.


Flos. 30-1 and 30-2. Distribution of
stresses in thick-walled cylinders for
a,r'
JJ

2» dr1 case «i — 0.

is equal to ira"p in the case of Eqs. (30-19) and equal to —irb"p when
Eqs. (30-20) are used. hollow cylinder having both ends closed which
A

stressed by an internal (or external) pressure


is

pulled (or compressed)


is
p

in axial direction just by these forces. Thus we see that the two groups
of equations (30-19) and (30-20) represent the solutions for hollow cylinders
having closed ends under internal and external pressure, respectively, and
we can add, since these formulas were derived by assuming that t, =
0,

MacGregor, Coffin, and Fischer in their paper quoted on p. 455 also arrived
1

at this special solution for finite deformation.


442 THEORY OF FLOW AND FRACTURE OF SOLIDS

that in both cases the cylinder does not change its length when it deforms
permanently. '
The ratio of the outer to the inner radius b'/a' when the axial length of
the cylinder remains constant («* = 0) according to Eq. (30-5),

decreases when the inner radius a' is increased. The internal resistance
of a cylinder having both ends closed when measured through the pressure
p [Eq. (30-21)] must therefore decrease as the logarithm of b'/a' decreases
when the radius a' is enlarged. The inner equilibrium of an ideally
plastic cylinder having closed ends yielding under internal pressure
according to the law expressed by Eq. (30-15) is therefore unstable if the
pressure p under which it is loaded is maintained, for example, at a con
stant value just required to make the cylinder yield through its entire
wall.
30-3. Radial and Axial Flow. In the general rotationally symmetric
distortion of a cylinder material elements will be displaced in both the
radial and axial directions, but the axial strain e, = to will not depend on
the variables r or tJ. To satisfy the condition of flow Eq. (30-15) valid
for an ideally plastic material it is noted that a plane ay = 0 intersects the
surface of yielding f(fr,<rt,<r*) = 0 which is a circular cylinder in a system
of rectangular coordinates oT,at,Ot in the ellipse

<r«2
- <rt<r, + <r,2 = <r02 (30-23)

and that the projection of any congruent ellipse on the o> = 0 plane, in
which a plane cy = const intersects the surface, is obtained by substitute
ing in Eq. (30-23) <r< — oy and a, — oy for at and <rz in each term :

(ft
~
*rY ~ (<T<
- Or){*z
- «r) + (f, ~ <Tr)2
= <T0S . (30-24)

This is a second convenient way to express generally the condition of


plasticity of Eq. (30-15). Now let
1 This has been well verified in
precise tests by E. A. Davis on the flow of heavy-
walled cylinders of steel and copper, mentioned in Chap. 17, p. 255. Since the state
of strain in a cylinder for which the axial strain vanishes (e« = 0) is a pure shear, we
may add that t, = 0 represents the condition of flow in cylinders having closed ends
for any general law of deformation valid, for example, in strain hardening ductile
metals (see Chap. 31). For this same reason P. W. Bridgman (J. Applied Phys.,
vol. 19, p. 302, March, 1948) found that hollow heavy-walled metal cylinders with
closed ends which he had exposed on their external surface to very high hydrostatic
pressure did not change their length appreciably. The metal was displaced radially
inward until the cavity was filled with the metal that had been displaced.
THICK-WALLED PLASTIC CYLINDERS 443

— =
a - a' »
at <rr t=—

a.
- ar =

(30-25)
V2
Eq. (30-24) is transformed into

+ 3<r" = W (30-26)
or
—^ + —= = 1 (30-27)

which in the new stress variables a and a' is the equation of these
'' ''
plasticity ellipses. 1
Treating first the flow in a cylinder when the strains remain infinitesimal,
the plastic stress-strain relations

/ at + a\ <t>

(30-28)

when combined with the expressions for and given in Eqs. (30-3)
tt

«r

show that a and must be equal to


a'

(30-29)

T
V2 /*„ ~
\
c


~
,

a
(
'

\2 W*)
i

and that the ratio

VZa' = \/3 - (2c/V3«0)(l/r2)


(30-30)
<r + (2cAo)(l/r2)
1

For the two unknown components of stress a and the two conditions
a'

(30-30) and (30-26) are thus prescribed. It may be convenient now to


change the independent variable by introducing an angle a defined by
r

tan a =
(30-31)
V3 *0r2

Equations (30-25) may be interpreted as a coordinate transformation consisting


1

in a parallel shift of the origin by an amount equal to a, in both directions at and<T,


combined with a rotation of the new axes a and by an angle of 45 deg.
a'
444 THEORY OF FLOW AND FRACTURE OF SOLIDS

Equation (30-30) may be expressed as follows:

- tan a
\/3
"
a' — V3 = tan
(H
- a (30-32)

I
^
V3

J
1 + tan a

and Eqs. (30-26) and (30-32) are satisfied if we take for

a -
\/2 fro cos — >
^
-
(30-33)
V3 = V2

<r'
<r„ sin

(g

«)
One more equation available, namely, the condition of equilibrium

is
of the stresses <r> and <r<
:

But the right side of Eq. (30-34) equal to (2<r0 sin a)/\/3 and after
expressing the left side of by using a is
instead of as variable, we obtain
it

r
- -^1cos a
p

=
(30-35)
da

as the differential equation which must be satisfied by the radial stress a,.
It has the integral

<To da <ro (r— + a\


J/ f

a, = =
,

In tan + ci
,

,
-= -= (30-36)
cos a o
y/Z
y4
(

2J)

ci being an integration constant. The tangential and the axial stress


are given by
2<r0 .
= sin a
H,

<r< <r> r=
V3
(30-37)
. /2x
\

2<r„
,

and thus the three components of stress ar,at,a, have been determined
in

function of the angle a which depends on the variable radius [Eq.


r

(30-31)].1
When either = or t0 = variable a cannot be defined by Eq.
a
0

0
c

(30-31); obviously by assuming = = u = —


«0/2) the trivial case
0,

(er
c

of a cylinder obtained which uniformly stressed by a system of stresses


is

is

Expressions similar to those in the text were developed in paper by the author,
1

Trans. ASME, 1929.


THICK-WALLED PLASTIC CYLINDERS 445

a, = a i, <r, = at ±
<to in the plastic state. When, on the other hand,
to = 0, the case of the cylinder without axial extension is obtained, which
was treated in Sec. 30-2.
Excluding these two cases («, = 0 and c = 0) in a hollow cylinder
having a and b for the inner and outer radius, according to Eq. (30-31)
there will correspond to r = a and r = b, respectively, a value aa and ab of
the variable angle a, for which

2c
a2 tan aa = b2 tan ab =
— — = const . (30-38)
7=
V3 «,

The integration constant d in Eq. (30-36) is determined from the bound


ary conditions for r = a or r = b.
For example, if the cylinder is exposed on the inner surface r = o to an
internal pressure p, we shall prescribe the boundary condition at the
other, free surface; r = b, a = a&, <r> = 0 and obtain for the radial stress
a, from Eq. (30-36) the expression

tan [(x/4) + («i,/2)] '


^3
and for the pressure

V3 tan [(t/4) + (o»/2)]


For on the contrary, under an external pressure p,
a tube yielding,
one has to interchange the angles aa and ab in these last two expressions.
It will be seen that the formulas (30-36) and (30-37) expressing the radial,
tangential, and axial components of stress, including Eq. (30-40) for the
pressure p and Eqs. (30-19) corresponding to the special case when t, = 0
have a perfectly analogous structure. Instead of the variable radius r
(when «, = 0) certain simple trigonometric functions express the depend
ence in corresponding terms on the variable angle a when the cylinder
deforms in both the radial and the axial directions.
If a < r < b, aa > a > a». The values of aa and a* according to
Eq. (30-38) depend on the ratio of the constants c and «0. We shall con
sider that value of aa or of <*6 as a characteristic parameter determining
the type of flow in a cylinder which corresponds to the free surface of it.
Thus, for example, in the case of internal pressure, when <r> = 0 for r = b,
a = ab, Eqs. (30-37) show that for
ab\

r = b , <r, = —£z sin a6 , a, = —yL sin — •


(30-41)

On the outer surface of the cylinder stresses at and a, are represented by


446 THEORY OF FLOW AND FRACTURE OF SOLIDS

sine curves in function of the angle with a phase difference of 2x/3.


Since in this case <r, > 0 for r = 6, the angle at can vary only between the
values 0 < a* < x. (a* = 0 cor
responds to simple axial tension,
ab = it to compression.) For
ir/2 we obtain the case t, = 0.
ab =
Conversely, if r < a, < 2x, assum
ing <r> = 0 for r = a, this furnishes
the cases of tubes under external
pressure, for which <rt < 0 for r = a.
Two examples are reproduced in
Figs. 30-3 and 30-4; the first rep
resents a tube pulled by a moder
ately large tensile force, the second
one compressed by fairly large com
pression stresses in axial direction,
while in both cases an external
Fio. 30-3. Fio. 30-4. pressure p acts on the surface r = 6.
Flos. 30-3 and 30-4. Distribution of stress When the strains increase to finite
in thick-walled cylinders subjected to exter values the preceding equations can
nal pressure.
not be used because Eqs. (30-12) are
valid neither for finite conventional nor for natural components of strain.
For Eqs. (30-12) stress-rate-of-strain relations must be substituted:

Ut
( a, + <rA
= - $2-
+
V■-—2-)
- 3a') (30-42)
(<r

U,
2 V2
which, together with Eqs. (30-11), permit the evaluation of the stresses
a and in function of the variable radius Since Eqs. (30-11) or
a'

r'

Ur - ~ 7*0
_ co
pi
'
2"

(30-43)
Uo Co
ut =
,

~l~
'

r'*
2

and the condition of equilibrium

,d<rr
—7=
1

-r-, = — a
(a

(30-44)
r

represent functions of the variable radius which have the same form
r'

as
THICK-WALLED PLASTIC CYLINDERS 447

the corresponding equations (30-34) had in function of the


(30-3) and
radius r when only infinitesimal strains were considered, we note that
the method for integrating Eq. (30-44) for a finite distortion in a cylinder
remains in all details the same as for infinitesimal strains. Instead of the

r',
quantities u, however, the symbols ur, u,, u0,
r, a, a', have to

b'
tT, «o, b,
be substituted. The corresponding expressions for the stresses, etc., are
thus readily obtained.
CHAPTER 31

FLOW IN METAL CYLINDERS CONSIDERING


STRAIN HARDENING
In the preceding chapter rotationally symmetric cases of flow have
been treated in hollow cylinders under the assumption that the yield
stress was not dependent on the amount of plastic deformation. We have
seen that the assumption of an ideally plastic substance has made it
possible to establish the exact solutions for such cases in finite form.
The theory may now be extended to the more general case in which, as in
the strain hardening ductile metals, the flow stresses in the material
increase with the permanent strains according to some law, which may be
known empirically or may be given through a mathematical expression.
Suppose that the behavior of a ductile metal is characterized by a
strain hardening function
r=/(7) (31-1)
where t denotes the octahedral shearing stress

r =
J V(»i
- *tY + (a.
- <r«)2 + (<rr
- a,)* (31-2)

and 7 the octahedral shear strain. If the strains are small,

7 =
| V(U
~
«r)2 + (e.
- £,)2 + («r
" «.)*. (31-3)

Since for incompressible material


er + €, + «.= 0 , (31-4)

7 may also be expressed; e.g., after eliminating tr, by

7 = 2 V% V«2 + «*. + ««2 • (31-5)

If the
strains are of finite magnitude, in Eqs. (31-3), (31-4), and (31-5)
instead of the conventional strains er,et,tt the natural strains lT,h,h must
be substituted and a unit shear y defined by

7 = 2 y/% vV + hi, + u2 . (31-6)

31-1. Theory for Small Strains. As long as the strains can be con
sidered as small quantities we may utilize various expressions which were
established in Chap. 30. If e. = «o = const according to Eqs. (30-3),
448
CYLINDER CONSIDERING STRAIN HARDENING 449

£r ~ _ «o
_ c
2 H
c
(31-7)
— «o .

Therefore

= 3*T
(31-8)
V2(e°2+£)
and from the stress-strain relations (30-28) in conjunction with Eqs.
(31-7) and (31-8)

«l

«r = K"t

*r) = Tj- V"
- "r) =
^ (31-9)

so that the stress difference


Act
=
<Tt <rr
yr* (31-10)

and from the condition of equilibrium

Act
r j—
do>
= <t<

<rr
= — , •
(31-11)
ar 7r2

After the variable r is expressed by means of y using Eq. (31-8)

£
=
-J vV - 2V (31-12)

from Eq. (31-11) the radial stress ar may be computed by a quadrature

IS
f rdy ,r (31-13)

where t =
/(?) and from Eqs. (30-28)

" = *r ^ ~
"\l

+
T 2t°2
'

vV -
(31-14)
2«„^

,r. =
^3eo
I

+ -
+

<rr
j

These last equations thus express the distribution of the stresses <rr,<Tt,at
in cylinders of metal which deform according to a strain hardening law
T = f(y) within the range of small strains.
1

Before the square root appearing in Eq. (31-8) and subsequently in (31-13), etc.,
1

a sign should be considered and the correct sign be determined in accordance with
±

the boundary condition for = a or = as the case may require.


r

6
r
450 THEORY OF FLOW AND FRACTURE OF SOLIDS

Considering, for example, the case of a hollow cylinder without axial


extension, *0 = 0 under internal pressure p since for r = b, ar = 0, the
radial stress <rr and the pressure p are expressed by the definite integrals

having the limits

/2
2c
/2 2c /2
2c ...

o-t and <r, are found equal to

ai = <rr + 2 a/M t
(31-17
0-.
= ffr + \/M T >

representing the distribution of the stresses in a cylinder having closed


ends of strain hardening metal under small strains.
So far none of the above developed expressions contain an assumption
with regard to the general stress-strain law t = f(y) under which a
monotonously increasing function of the shear strain 7 was postulated.
31-2. Special Cases of Cylinders with Closed Ends. Suppose that
t = f{y) is a power function of 7. The two integrals in Eq. (31-15) may
for example, the shearing stress
if,

be evaluated easily a power

is
r
function
= tot" < m <
(0

(31-18)
1)
r

of the shear 7,1 in which case one finds from Eq. (31-15) that the radial
stress expressed by
is

one introduces the radial displacement u = ret = c/r, one can define
If

the constant in Eqs. (31-7) by = uaa, by means of the displacement


c

u0 at the inner surface


= o of the cylinder in which the pressure
is
p
r

transmitted. The shear then also expressed by


is
7

uaa = a2 . „
/2

=
\3 7"
7

"r* r*
(

as a function of r. This makes possible to express also as a function


it

oy

of through
r

Ta a la
\
I
\

See author's paper in J.


Applied Phys., vol. power function has
8,
1

p. 418, . 937.
A

been also considered inpaper by A. Winzer and W. Pbaoeb (J. Applied .\fechanirt.
a

preprint A-6, 1947, ASME meeting).


CYLINDER CONSIDERING STRAIN HARDENING 451

and therefore the pressure

where t„ is the shear stress ta = T0yam corresponding to the shear

With the abbreviations p = r/a and a =


a/b the three components of
stress a„ai,a, in a cylinder having closed ends under internal pressure p

3.0f\-

I.Of
CruNXR WALL

-O.Sp\

SHEAR f -

Fio. 31-1. Strain hardening function Fio. 31-2. Stress distribution in a heavy-
t raym. walled cylinder when t = T07m, (a/b =

for a material yielding according to the strain hardening law t = Toy"


may be expressed by the functions as follows :

-
V 1
= — *»
<rv

-
1

_
a2"j

V (1 2m)
„ -
(31-23)
— or
-
1 p2m
,j

- (1 m)
a2mj

+
1

This distribution of the stresses has been illustrated in Figs. 31-1 and
31-2 assuming for the exponent m = J^, %, and As special
0,

1.

Y±,
452 THEORY OF FLOW AND FRACTURE OF SOLIDS

cases of the power-function law, when m = 0 the case of an ideally plastic


substance yielding under a constant yield stress t0 = const and when
m = 1 the elastic substance is obtained. We note from Fig. 31-2 when
the exponent m = J^ that the tangential stress <rt is constant throughout
the entire cylinder. When 0 < m < the greatest value of <rt occurs

]4,
on the outer and when on the inner surface of the cylinder.1

Yi
< m <

1
a. Plastic Flow in Moderately Thick-walled Cylinders. Equations
(30-19) and (30-20) show that the difference of the tangential stresses at

taken between their values on the outer and on the inner surface

of
a
heavy-walled cylinder of ideally plastic material

(*)*•
- («r«).'
= ±P (31-24)

just equal to the absolute value of the pressure p. This also true
is

is
for any heavy-walled cylinder of an elastic material (the case m =

in
1
Sec. 31-2) for any finite ratio of b/a. Thus the value of the tangential
stress which the largest of the stresses drops in heavy-walled cylinder
is

a
from its greatest to its smallest value in an elastic and in an ideally plastic
cylinder by an amount equal to the pressure p.
Based on this fact one may construct an approximate representation
of the distribution of the stresses <rt and <rr in moderately thick-walled
cylinders having closed ends or in cylinders exposed to internal pressure
in which the thickness of wall = — a not too large fraction

of
is
p

a
h

the mean radius r0 = -\- b)/2 assuming that the strain hardening func
(a

tion = To7m. The curves representing the stresses ar and o-< as function

a
r

of the radius may then be taken approximately as straight lines across


the wall2 and consequently the mean tangential stress o-(0 in the wall

is
equal to

When m = some of the formulas fail, but in this case, by direct computation
1

Eqs. (30-19) and when to = the Lame formulas for an elastic cylinder are obtained;
1

a heavy perfectly viscous substance would be deformed in the same manner. One
of

verifies also easily that the resultant force exerted in axial direction, independently
the value of the exponent m, equal to t(6* — a*)p so that the formulas hold for
is

cylinder having closed ends in which the state of strain at every point a pure shear.
is

Since the same relations hold (when e, = for a material in which the shear stress
of y 0)

t a function of the rate of shear dy/dt = and in which = f(y) may express the
is

dependence the yield stress on small rates shear y, Eqs. (31-7) to (31-9) are also
of

valid i„it,i, are substituted in them instead of er,«i,«.. We conclude that all expres
if

sions may be taken for expressing the corresponding cases of the slow creep in metal
cylinders at elevated temperatures following a general creep-speed latv t = f(y) including
the special cases when the creep-speed law expressed by power function «■rW"
is

This equivalent to stating that in such cylinders the expressions containing the
2

is

ratios a = a/b and r/b may be developed in powers of the small ratios a " h/r, and
— r„)/r<, by neglecting the powers higher than the first of these quantities.
(r
CYLINDER CONSIDERING STRAIN HARDENING 453

pa ( r0 l\
(31-25)

After denoting by x = r - r0, a = r0 - (h/2), b = r0 + (ft/2) where


the wall thickness h = b — a, from Eq. (31-23) or

after developing a2m = (a/6)2m and p =


r/a in powers of the small
quantities h/r0 and x/r0, we obtain /or the distribution of the tangential
stress <rt across the wall the formula
/
J?
ELASTIC nc
PLASTIC\
[h [>„

" ~
1

= + 2m)
(1

T
V

,
2

I
(31-27)
which may also be written as

°t = + - 2m) H (31-28)
(1

<rt0
/i

and for <Ae radial stress

* = -p - •
(3i-29)
1)
G

The slope under which the straight


line representing the stresses at in
is

clined — 2m)p/h. Thus we note


is

Fio. 31-3.
(1

Fio. 31-4.
again, when m = that the slope Fioe. 31-3 and 31-4. Moderately thick-
is

walled cylinders.
zero and when m }4 that posi
it
is
^

tive and negative, respectively. The two distributions of stress for the
exponents m = and m = are represented in Figs. 31-3 and 31-4.
0
1

Cylinder with Closed Ends Yielding under General Strain Hardening


b.

Law at Finite Strains. Let r,a,b be initial values and r',a',b' the corre
sponding values in the strained condition of the cylinder. If the axial
strain vanishes, e* = u, = =
i,
0,

we have
0,

- - - a'' -
r/'

a'' = r2 a2 and 6'' = a2


62 (31-30)
and from the stress-rate-of-strain relation

w, =
*(^-^p) = o, V, =
fr + <r<
(31-31)

the octahedral shearing stress t


is

= ~
sVI0" *r)
7

(31-32)
454 THEORY OF FLOW AND FRACTURE OF SOLIDS

where the positive sign was taken assuming that the cylinder is stressed
under an internal pressure. Since in the case when «, = 0 the state of
strain in every material element is a finite pure shear, the natural
octahedral unit shear, on account of er = — e<, i, = 0 is

7 = 2>/r; =
(31-33)
>/Iln£"
From the condition of equilibrium:

r = <r'-^ = (31-34)
57 3V3T
we obtain the radial stress o> equal to

~ 3
2 l b'Tdr'
/
— •
(31-35)
A/3

The integrand may be expressed by using y as the variable. Since,


according to Eqs. (31-30) and (31-33),

r* _ a'' - a2 —
16 -"Vf-, 2dr'— J3 dy
1 .9 .J
\2 -
S

r>

aS
r>'

'

(e^2

1)
(31-36)

we may express the radial stress and the pressure also as follows
p
<rr

fy rdy rdy
f'
3

=
(31-37)

(c,2'-l) "
(e^T-l)
The limits in these integrals are to be taken

y, 7a, 76 = In (31-38)
,

respectively. The stresses <rt and a, are given by

a, = + 2aJt, =
+

<rr <r, <rr (31-39)


r
.

For small values of the expressions derived in Eqs. (31-37) converge


y

to the ones which were formerly obtained for small strains in Eqs. (31-15).
1

theory of the more general case of combined distortion of a heavy-walled cylin


1
A

der in radial and axial direction sustaining finite strains and assuming a strain har
dening function for the metal has been treated in the author's paper in volume
a

commemorating the seventy-fifth birthday of Hans Reissner, New York, published


CYLINDER CONSIDERING STRAIN HARDENING 455

c. Instability of Equilibrium Causing Local Bulging in Plastically Deform


ing Cylinder. From experiments on the plastic flow of heavy-walled
long cylinders of ductile metals stressed under internal high pressures it is
known that after the pressure reaches a maximum the cylinders cease to
deform uniformly but bulge out locally before they finally burst. The
condition of the instability of the plastic equilibrium in cylinders may be
expressed generally as follows. As a convenient measure of the distor
tion suppose that a variable parameter q = a"/a* is introduced. The
strain at the inner surface r' = a' of the cylinder being equal to

* = ln = ln q •
a 2

we have q = e2'1 and always q ^ If


furthermore /3 denotes the con
1.

stant ratio 0 = b2/a2 where a and b are the inner and outer radius of the

cylinder in its original unstrained condition, according to the Eqs. (31-30)


we have

^
= q ,
^
= 1
- /J + fiq . (31-40)

The two limits of the definite integral Eq. (31-37) expressing the pres
sure, according to Eqs. (31-38) are seen to depend on the parameter q as
follows:
7a = VMln?, y> = V% In (1 - /S + (Iq) (31-41)

while q does not appear within the integrand itself and therefore, by
partial differentiation of the definite integral with respect to the param
eter q on which both limits depend, one obtains for the rate of increase
of the pressure with q:

dp 3 Ta dya n dyb
(31-42)
2 ^y
l(e Vz - ■1)
dq dq dq
,
(e
> •2 r
_ •1)
After substituting here for the derivatives of the limits ya and 76 their
values from Eqs. (31-41) and for

by Edwards Bros., Inc., Ann Arbor, Mich., 1949. C. W. MacGregob, L. F. Coffin,


and J. C. Fischer, The Plastic Flow of Thick-walled Tubes with Large Strains (J.
Applied Phys., vol. 9, No. 3, pp. 291-297, 1947), developed equations for the same case
considering also finite strains. In their analysis they assumed stress-strain relations
according to Eqs. (30-12) in which the natural strains ir,h,h were utilized. Since the
basic stress-strain relations were assumed by them in different form from the one
postulated in the first mentioned paper, results of both computations cannot be com
pared, with the exception of the case when the axial strain t, vanishes, in which case
both gave the same result.
456 THEORY OF FLOW AND FRACTURE OF SOLIDS

e
y2y° =
q , e
>2" = 1 -p+ jSg (31-43)
one finds

g- 1-P

Vff
dg + to)

(
1

If tends to reach an analytic maximum during the dis


the pressure

p
tortion, by letting dp/dq = the latter condition should make

it
0,
possible
to compute the value of at which becomes a maximum. If the root

p
q
satisfying the condition dp/dq = real, this defines the instant at which

is
0
local bulging must begin. For an extremum of we therefore have the

p
condition
- r^frr, <3'->5>

'-;
provided that = does not occur simultaneously, in which case obvi
1
q

ously dp/dq in Eq. (31-44) would become 0/0.


F. Koerber, C. W. MacGregor,1 and others have observed that the
tensile stress-strain curves of most ductile metals = f(y) beyond the

r
load maximum tend to approach straight lines. Suppose that the entire
strain hardening curve
=
/Or) = t0(1 + cy) (31-46)
t

consists of straight line, the octahedral shearing stress referred to


is
t
a

the actual dimensions in the minimum cross section of tensile bar or

in
y a
an element of the cylinder in its strained condition and the correspond
ing natural octahedral shear strain; to and are material constants.

If
c

the actual curve = f(y) considerably deviates from a straight line, as

is
r

usually observed for values smaller than the shear corresponding to the
y

tensile load maximum, Eq. (31-46) may represent the straight line which
observed beyond the load maximum.
is

At the beginning of complete yielding, when = only very narrow


a
1
q

band of values of along the stress-strain curve Eq. (31-46) near the
y

=
(y

ordinate t0 will represent the strains and stresses in the interior


0)

of the cylinder. Consequently this will start to yield under pressure


a

equal to

= =
2.|r„ln^ ^lln^ (31-47)
p

of cylinder of an ideally plastic substance [Eq. (30-21)]. Since


a

t„ = t0(1 = r0(l + y/%


-
cya) In
q)
+ +

^j^g,
,

I |

n = cyb) = V% In + 0q)]
+

t.(1 r„[l
p
(1
C

The Tension Teat, Proc. ASTM, June, 1940. MacGreqor and L. E. Welch,
1

AIME Tech. Pub. No. 1507, 1942; No. 1393, 1941.

X\
CYLINDER CONSIDERING STRAIN HARDENING 457

when these expressions are substituted in the bracketed term in Eq.


(31-44), the right side of the equation becomes indeterminate (0/0) for
the value of q = 1. After computing the true value of it for the limes
q = 1, we find the rate of increase of the pressure at the beginning of
complete yielding:

(fL-J^-wWI--1). <3«9>

This has a positive value when the constant c measuring the slope of
the straight line in the stress-strain curve Eq. (31-46) is

c > V% = 1.225 . (31-50)

Corresponding to a slope c = 1.225, yielding will start in the cylinder


under a minimum of the pressure p. When c < 1.225 the pressure will
drop when the cylinder starts to expand radially and its equilibrium
remains unstable. Only when c > 1.225 can a positive rate of increase
of the pressure develop followed by a maximum of p and a stable range of
expansion exist. At the pressure maximum from the condition

^ = 0
'
dq

from Eq. (31-44), using Eqs. (31-48) we obtain


+
-+ +
1 c,ya — 1 cyb
'
(31-51)
q 1 0 pq

If the values of ya and 7;, are taken from Eqs. (31-41) and substituted
in Eqs. (31-51) a transcendental equation for the unknown q = a"/*1*
is obtained. The root of this equation determines the value of the ratio
a' fa or the unknown radius a' at which the pressure p becomes a maxi
mum and the heavy-walled cylinders must start to bulge out locally.
CHAPTER 32

PARTIAL YIELDING IN CYLINDERS


32-1. First Yielding in Cylinder under Internal Pressure. For an

elastic material the radial, tangential, and axial stresses in a thick-walled


cylinder of inner radius a and outer radius b for the case of an internal
pressure p according to the theory of elasticity (Lame" formulas) are
given by

n—
— pa2
aT —
/'
"A
a2V2

+ <3i"

')■
"t

i
_
'• ~
2ypa2
+ Ee-
W^Ttf
>
where t, the axial strain in the cylinder, Poisson's ratio, and
is

E
the v
modulus of elasticity. Let the ratio of inner to outer radius be denoted

by
a = a/b and the yield stress in tension by a0; the condition of yielding
for an ideally plastic substance expressed by
is

(<r«
- <rr)2 + (it, - <rty + fa - a,)* = 2cr02 . (32-2)

The cylinder will start to yield at the inner surface = a. After sub
r

stituting the values of o>, <rt, and at taken from Eq. (32-1) for =

in
r

a
Eq. (32-2) we obtain the equation

+

3

£00" -r?MiD(*) +©'-'•


(1

For the two variable ratios p/a0 and <r,,/<ro this the equation of an
is

ellipse or for the unknown ratio p/c0 at a given ratio at/aa a quadratic
equation, from which the pressure under which the cylinder starts to
p

yield can be computed

p_
M*)*^-- -»(*)]
1

«•">
(r0 2(3 + a4)

the ratio cri/Vo known.


is
if

These are two solutions according to the sign taken before the square
If,

root. for example, az = for the upper sign we obtain pressure


a

<n>,
458
PARTIAL YIELDING IN CYLINDERS 459


p = (1 a4) and for the lower sign the pressure p = 0 (the
a2)aV0/(3 +
case of axial tension at = <ro, p = 0 in a cylinder).

For a cylinder without axial elastic extension (t. = 0),

-
=
2vpa*
Oz
1 a2

the pressure must be equal to


- «>■
-
(1
(32-5)
V3 + (1 2i>)2a«

A cylinder without axial stress {<r, = 0) or a flat ring starts to yield at

-
a pressure
(1 <*2>o

A cylinder having both ends closed within the elastic range of strains
will first expand slightly elastically by an amount equal to

= (1
- 2v)pai
«x —
\
1 a'2

in the axial direction. The axial stress a, being equal to

Oi —
1 - a2'
a2p

the cylinder will start to yield at the inner surface r = a at a pressure


equal to

P =
^p° •
(32-7)

32-2. Comparison of Three Cases. It may be instructive to compare


the distribution of the stresses in three cases which have been reproduced
in Figs. 32-1, 32-2, and 32-3. The first two refer to the case of a cylinder
with closed ends stressed by internal pressure having a fairly large ratio
a = a/b = 0.4. Figure 32-1 shows the stresses in an elastic cylinder at
a pressure p = 0.485ff0 [computed from Eq. (32-7)] when yielding just

starts, and Fig. 32-2 when the cylinder yields completely at a pressure
p = (2cr0/\/3) In b/a = 1.057<r0 [computed by using Eq. (30-21)]. The
third Fig. 32-3 refers to the case of a flat ring treated later in Chap. 33
yielding completely under an internal pressure p = o-0. It is noted again
in accordance with what has been found for cylinders yielding under a
strain hardening law expressed by a power function (Chap. 31) that the
460 THEORY OF FLOW AND FRACTURE OF SOLIDS

peak of the curve at the inner surface r = a in an elastic cylinder is


at
entirely wiped out after the tube yields completely throughout its wall.
32-3. Partial Yielding of Ideally Plastic Cylinder in Plane Strain.
When a cylinder of mild steel having a well-defined yield point is gradually
strained by a radial pressure and an axial load the strains in a certain

Fio. 32-1. Fig. 32-2. Fig. 32-3.


Fios. 32-1, 32-2, and 32-3. Distribution of stresses in thick-walled cylinders and a flat
ring subjected to high internal pressure. Radial, tangential, and axial stresses a„ a. o,.
Fig. 1, when cylinder is still elastic and just starts to yield. Fig. 2, when it is entirely
in the plastic state. Fig. 3, stresses in a flat ring in the plastic state.

portion of the cylinder will be elastic while in the yielded portion they
will consist of two parts and will be the sum of an elastic and a permanent
strain. For comparatively larger values of the permanent portion of the
strain Poisson's ratio v for the total strains approaches the value v = \4
of incompressible material. In the elastic region, however, v = 0.3 for
steel. To avoid the difficulty resulting from the necessity of considering
a variable Poisson's ratio increasing gradually from the value 0.3 to 0.5
within the plastic region of the cylinder, it is first suggested that the
constant value of 3^ for v be introduced in both regions. This amounts
PARTIAL YIELDING IN CYLINDERS 461

to assuming incompressibility both in the elastic and in the plastic region.


Considering the partial yielding in a cylinder under an internal pressure p
when the plastic zone has reached a radius r = c the stresses in the
elastic zone are given by the formulas which are derived in the theory of
elasticity:

, -
(. % .

for c ^ r ^ b (32-8)

<r, = Ci + cu ,

and the axial strain is


— +
'* "_ <t, {{ar <rt)/2] _ Co
= <0 • (32-9)
E E
When the plastic zone grows, new material elements will become gradu
ally plastic while in the material elements which have already yielded the
components of principal stress will change their values continuously.
Since the increments of strain in a given element are produced under
successively changing values of the three principal stresses satisfying the
condition of an ideally plastic material, within the plastic zone stress-
rate-of-strain relations should be considered for the permanent portions
of the strains [such as Eqs. (30-13) which were introduced for the states
of finite strain, but which would also be valid for small strains]. As a
total strain « is the sum of an elastic («') and a permanent strain (e"):
c = t' + (", when the rates of strain are computed i = i' + i" the terms
e' will contain the derivatives of the components of stress with respect to

the time / and the terms i" will depend on the stresses themselves. This
introduces considerable difficulty.1
The analysis of the process of partial yielding may be simplified by
introducing in the plastic region a ^ r ^ c instead of Eqs. (30-13) stress-
strain relations of the form of Eqs. (30-12) using them for the total strains
and letting Poisson's ratio be v = everywhere in the cylinder.
J-£

Assuming small strains and since the axial strain tz = «0 = const the
normal stress az may be expressed through ar, at, and eo from the last of

R. Hill, E. H. Lee, and J. Tupper pointed out a rigorous method for analyzing
S.
1

the problem of partial yielding in a cylinder of material having a well-defined yield


point (Proc. Roy. Soc. London, ser. A, vol. 191, p. 278, No. 1026, Nov. 18, 1947) and
completely carried out the synthesis of small elastic and permanent strains leading in
successive steps of increments of pressure to the distribution of the stresses <sT,at,a, in
the cylinder. The analysis cannot be reported hero because of its length, and the
reader particularly referred to this paper.
is
402 THEORY OF FLOW AND FRACTURE OF SOLIDS

the following three equations. In the plastic region we have to satisfy


the relations
30 , «
£0
2

3* , to
<rr)
2 (32-10)*

0> + 0"(\
(0>
Denoting by
— Or
= 0"« +=—
Cr ;
= 0"(
(T 1 ff 7=— (32-11)
V2 V2
; have
30 , 6q \

(32-12)
30 , «0 \

and from the condition of compatibility of the strains

(32-13)

£2
0<t'
r2 (32-14)

where c2 is an integration constant. The strains «r and et are therefore


also equal to

"
3c2
2
2V2rs
(32-15)
_ 3c2
*
«o ,

2+2V2r=
The constants Co,Ci,c« are determined from the boundary conditions.
The variable stress a' can be determined from the condition of plasticity
Eq. (32-2) after noting that, according to the last of the three plastic
*
As stated in Sees. 28-3 and 28-7 these relations are valid only if the ratios cf'i
and az/<rt remain constant in any material element during its permanent deformation,
which is not the case during partial yielding; however, we shall utilize them in this
form with the desire not to complicate the further computations in Sees. 32-3 and
32-4 unduly and realizing that the solutions thus derived violate the premises on
which Eqs. (32-10) were based.
PARTIAL YIELDING IN CYLINDERS 463

stress-strain relations, a, = -?—= —- + -j = —-= + -j so that from Eq.


* <P v2 <P

(32-2)

3"" + %' = W • (32-16)

Let k denote the constant

k-4t (32-i7)

From Eq. (32-16), using (32-14) and (32-17), a' is found equal to

• - I?
\3 yT+ (32-18)
ifcv

(the positive sign before the square root is taken since a' > 0 for the
cylinder subjected to an internal pressure) and consequently the flow
function <£ is equal to

The constant fc may now be determined


from the condition that on the

<f>
boundary surface of the plastic zone r function = c the flow
must
= l/E, where
<t>

become equal to the modulus of elasticity of the


E
is

material. This gives

= xln ~ ~ , and c* = c2 xj-= (-^ — e02


k

(32-20)

Furthermore, after equating the two values, which may be computed


for the stress from the expressions valid in the elastic and in the plastic
a'

region along the boundary surface of the yielded zone = by using


c,
r

Eqs. (32-8) and (32-18)

\3Vl
c2

+ )fcV

we see that Ci must be made equal to

-
*


Cl = -^= (32-22)
VI
,

V3
1

&2 **c«

The constants Ci,Ct,fc have thus been determined.


464 THEORY OF FLOW AND FRACTURE OF SOLIDS

From the condition of equilibrium

r
^
= a, - ar (32-23)

after expressing ar and <rf by means of a and a' using Eqs. (32-1 1) and after
substituting for a' its value computed in Eq. (32-18) since

Ydr
= =
It (rV) (32-24)
-vl^?^
T

the unknown stress a found after integration equal to

is
" "
(vi

*h)
** + + c,
>|
VI

2

"

(32"25)
+

/
+

r
Since the first term on the right side of the last equation just equal to

is
-
the radial stress <r> expressed by
is
a'

ar = = ^ /" 4=

+
(32-26)
V2 V3J[ VTTW( V2
or by r

* --
+
VI ^j- (32-20

The constant determined by equating this value of <r> taken for


is

c3
= and the corresponding value of <r> in the elastic zone
r

°r = -
^ •
(

cx
1

The radial stress in the plastic zone a of the hollow cylinder

is
^

^
r

expressed by the function

„ - * - A^Snil
-
P

= m + In
S

Vi *v vi (32-28)
2
7

lj

V3 + *v«
+
r
L

which also furnishes the value of the pressure under which the plastic
p

zone has advanced to the radius =


r

VHvttw4 vr+W4-iJ
a

The tangential and axial stresses in the plastic zone


g

a are
^
r

<r<
= =+ <Tr
.

(32-30)
a0kr2 a, + at

Vl
^_
+ fc2r4
PARTIAL YIELDING IN CYLINDERS 465

It will be noted according to Eq. (32-20) that partial yielding occurs


unless the constant k has a real value or unless — <r0 < toE < <r0. (If
Et o = ao, k = oo , the axial strain e0 has just reached the value correspond
ing to complete yielding by simple tension a, = co = const.)
32-4. The Partial Yielding in Cylinder Having Closed Ends under an
Internal Pressure p. Suppose that Et0 is small compared with «r0.
Then he* becomes small compared with 1 and the last term on the right
side of Eq. (32-29) becomes (assuming fc2a4 and A!c4 as small quantities) :

VI +- k2cl
- -
VTTW4 -
1
In = In = 4 In (32-31)
l
Thus we see that, when the axial strain «0 vanishes, the pressure p
and the radial stress ar in the plastic region a < r < c are expressed by the
simpler formulas

p =
V3\(
1^ i _ Is +
b1
2 In £
a)
V (32-32)

ar - - -^= (l - £- + 2 In i\, (32-33)


V3 V V r)
while the stresses a, and ot become equal to
ln0

V3\ *>2

,. - .£• fe - (32-34)
2

V3V>2
In the elastic region the stresses are
^

^
6
r
c

V3 62 \r2 V3 62 \r2
/

,,-=-£*£!. (32-35)
V3b2
Equations (32-32) to (32-35) express the distribution of the stresses
in the case of partial yielding in a cylinder of ideally plastic material
having both ends closed and yielding under an internal pressure p.
Equation (32-32) indicates the value of the pressure when the plastic
p

zone has advanced to a cylinder of radius c.1


In Fig. 32-4 the distribution of the radial, tangential, and axial stresses
in a hollow cylinder shown for the ratio of inner to outer radius
is

a = r = 0.4

These formulas were given in the first edition of this book. Figure 32-4 contains
1

certain corrections which were added later.


466 THEORY OF FLOW AND FRACTURE OF SOLIDS

and for four different values of the pressure p corresponding to a plastic


region which penetrated to c/b = 0.4 (first yielding), 0.6, 0.8, and 1
(complete yielding).

0
<^aIO.8
2.,,
0>.<?,<* __OJ
1 1*' 0.6
1.6
\-oa 05

. 04
•0.6
z \
a8
TANG.B,
j0.4 ''AM . 03

J
,.4
6~~—
.,4
as
AXIAL a

/<06s
-0.4\ 4x/3

-,.8
W

//•o.a
I'
Ran st} lESSaj. *--I.O
0 .2 .4 .6 8 1
-IS
a -
« c
-1.6 1
.>- b
PLASTIC ELASTIC
-2,L
O ,.2 ,.4 Q6 ,.8 10

^^^^^^
-r-t>-
Fig. 32-4. Partial yielding in thick-walled Fio. 32-5. Partial yielding in tubes.
tube subjected to internal pressure p. Curves indicate how pressures p would
Curves indicate distribution of radial, tan increase in thick-walled tubes with outer
gential, and axial stresses <r„ at, <ri when radius c of plastic region. The pressure
c:b = 1, 0.8, 0.6, 0.4; a inner, b outer, r vari curves are plotted for 10 different ratios of
able radius, c radius of outer boundary of or = a/6 of inner to outer radius of tubes.
plastic region. (For Poisson's ratio „ = ^ is taken.)

In Fig. 32-5 the internal pressures p required to produce progres


a second
sive yielding in hollow cylinders having closed ends are plotted downward
for eight different ratios of a = a/b = 0.1, • • • ,0.8 above the values
of the ratio y = c/b, thus indicating the increase of pressure p with the
spreading of the plastic region.

The case of partial yielding in a thick-walled cylinder of a metal exposed to any com
bination of internal and external pressure and an axial load has been investigated by
PARTIAL YIELDING IN CYLINDERS 467

C. W. MacGregor, L. F. Coffin, and J. C. Fischer,1 who assumed that the metal has a
well-defined point and after this has been reached deforms according to a strain hard
ening law postulating that the octahedral shearing stress t0 is a known function
T° = /(7o) of the octahedral unit shears 70 which they assumed as small shears. Since
in their computation they used stress-strain relations in a form identically correspond
ing to Eqs. (32-10), the same remarks are in order which were made in the footnote
with reference to Eqs. (32-10). They determined by numerical methods the distribu
tion of the stresses a„ <rt, and a. in tubes of various sizes whose metal complies with
the laws of yielding under a constant value of the stress (to = const) treated in this
section.'
32-6. Spreading of Plastic Zone around a Cylindrical Cavity in Which the Pressure
Is Increased. In the case of partial yielding, which is discussed in this section, the
variation of Poisson's ratio within the growing plastic zone and the elastic compressi
bility of the material are considered. As in the preceding section and with the same
restrictions the assumption will be made that within the plastic region stress-strain
relations hold in the form which is valid for small strains when the ratios of the princi
pal stresses in a material element remain constant. These relations will be expressed
for the total strains. For example, the total strain t, is the sum of an elastic («,') and
a permanent strain («,") :
tr = *,' + «r" . (32-36)
For t/ we have
tr' = <t>'[<rr - y'(o, + <r.)] (32-37)
<t>'

<t>'

where we write for = 1/2?. For er" we have

tr" = 4>"{<rr
- v"(*< + <r.)] (32-38)

where <t>" variable quantity. For most applications we may assume that /' = J^.
is
a

The total strain may therefore be expressed by

* = 4>[<r, -
+

x(<r, <r,)] . (32-39)


<t>

and denote [See Eq. (24-1 la)]


v

= + r-^w+y'- (32-40)*
*
<t>

The flow function and the quantity representing a variable Poisson's ratio in the
v

plastic zone are functions of r. varies between the values 0.3 ~ < < t/' — 0.5.
*'

r
v

"Partially Plastic Thick-walled Tubes," Franklin Inst., vol. 245, No. pp. 135-
1

2,

358, February, 1948.


They evaluated the components of stress in metal cylinders in which
2

plastic
a

zone develops under the action of an internal pressure and in which the resultant force
in the axial direction of the tube vanishes. Among previous attempts at dealing with
similar cases of partial yielding, which they quoted, might be mentioned those by A. C.
Macrea, "Overstrain of Metals" (His Majesty's Stationery Office, London, 1930)
and by C. F. Jeansen, "A Treatise on Radial Expansion of Guns," Navy Department,
Washington, D.C., 1938 (utilizing a flow condition related to that of Coulomb,
atrributed to Duguet). On partially yielded metal cylinders see also paper by D.
a

N. de G. Allen and C. R. Sopwith, Proc. 7th Intern. Cong. Applied Mechanics,


London, September, 1948.
* The Flow
of Metals under Various Stress Conditions, Proc. Inst. Mech. Engrs.
London, vol. 157, p. 121, 1947.
468 THEORY OF FLOW AND FRACTURE OF SOLIDS

We have, furthermore, the identity

(1

2^)0 = (1
- 2/)*' = k = 1 3/C . (32-41)

where K = 2?/3(l — 2r') is the hulk modulus of the metal.


Consider the spreading of an advancing plastic zone around a cylindrical cavity
of radius a around which the metal yields under the action of an internal pressure
p while the axial strain t, = 0. In the elastic portion of the body c < r < « the
stresses are

°r - —oi = ~t
, a. = 0 . (32-42)
^7=

The stresses or,<n,a, in the boundary surface r = c between the elastic and the plastic
zone satisfy the plasticity condition Eq. (32-2).
In the plastic zone a < r < c the components of stress a^ai.a, must satisfy the plastic
stress-strain relations. These, according to Eqs. (32-39) for the total strains t„n,u
take the form

«r -^- (1 + —
a.) ,
*

" = (1 + v)<t>(*, — a,) , (32-43)


C —

= 0 , a, = v(<r, + "<) •

With these relations and the transformations [Eqs. (32-11)]

"T" ff> , &t "~ <Jt

the plasticity condition Eq. (32-2) now takes the form

(1
- 2»)V, + 3o'J = 2<r„2 . (32-45)

In this ellipse the major semiaxis v/2o-i0/(l—


2„) increases indefinitely when r
approaches the value v = The major axis has its smallest value \. 2 "o/O — 2/)
when » = = 0.3, and the larger the permanent strains are the longer the ellipse
!''

becomes.
<r'
The components of stress <r and a' when yielding just starts define the point >I

y/%oo, a = at the end of the minor axis of the ellipses Eq. (32-65). The points rep
0

resenting the plastic states will be crowded near this region but will belong to different
ellipses corresponding to the changing values of and as v approaches the value the
^
v

locus of the points {a, a') approaches the common tangent to the ellipses in point "
«'

-\Z% <ro, a = 0. In first approximation the points a,a' will lie on this straight line

const
.

To satisfy Eq. (32-45), however, we shall assume that

V2 1to sin
0

— 2„
1

(32-40

-4 cto cos
/3
.
3
PARTIAL YIELDING IN CYLINDERS 469

The angle /3 introduced in these equations is equal to zero in the boundary surface
t = c of the advancing plastic front.
The condition of equilibrium in the plastic region

-<r,
r^ = <r, (32-47)

in terms of the variables a,a' is expressed by

r£—dJrV)=0
dr r dr

If x denotes x = r2, after substituting the values of a and a' from Eqs. (32-46), Eq.
(32-48) is also expressed by

•a(£i)+7i=C"-»-0. (32"49)

After making use of the stress-strain relations Eqs. (32-43) the compatibility condi
tion Eq. (32-13) takes the form

d
sin - 1

d
-
fji

x (z/i cos

0)
«

(32-50)

0
w here m stands as an abbreviation for

+
y
1

(32-51)

If
1/(1 - 2,) =
in Eq. (32-49) 1/(1 — 2v) by its value in terms of n, namely, by
we replace
+ 2M)/3 we have two differential equations [Eqs. (32-49) and (32-50)]
i
(1

to determine the two variables and in function of = r*.


m
0

It appears unnecessary to compute exact integrals of these two equations, even

if
<r'

this were possible. The stress a' being practically equal to = *o = const,
Eq. (32-48) shows that a must approximately satisfy the equation

=
r^ \^o (32-52)

and that a and a' may therefore be expressed by

ct = „ sin = —
\5o <to In
-
/3

—7— — £v
*

r
1

(32-53)

videos, =
vi-. vi—^^Hy,
where the angle remains small in the entire interval a < < so that also
c,
/3

p.zo^M^f. (32-54)

The second term under the square root in the expression of a' [Eq. (32-53)] vanishes
for = and becomes small when v approaches the value As the numerical evalu
c
r

ation shows, the stress o' expressed with sufficient accuracy by


is
470 THEORY OF FLOW AND FRACTURE OF SOLIDS

, + I2 r, 2(i -2*)*/. c\n (32-55)

and Eqs. (32-44) and the last of Eqs. (32-43) therefore make it possible to evaluate
the components of stress in the plastic zone a < r < c:

[- - -
1 2 In - + 2(1 2k)
V3 ('»;)■]•

(32-56)

a. - -^ \ - 2 In-r + - 2k) In -1 •
V3 L
2(1
r J
The unknown value of Poisson's ratio v appearing in the second terms on the right
sides of these three expressions may finally be computed approximately from the

-ELASTIC REGION

Fiq. 32-6. Partial yielding around a cylindrical cavity under internal pressure p.

compatibility condition Eq. (32-50) after noting that the first term in it containing
sin /3 is quite small compared with the second term. After neglecting the first term
and after writing for cos 0 = 1 in the second term, n may be computed from

rf(XM) = ,c«
0, (32-57)
dx

denoting by the value of for = or = "')/(! — 2k'). Consequently


pi

+
/»'

(1
c
r
it

the variable Poisson's ratio k approximately expressed by the function


is

- -
+ y')c*
- 2KQr' £r
(1 (1
(1

£c)
(a

(32-58)
2(1 + r')c* + 2k>«

The pressure under which the yielded zone has advanced to depth = given
is
p

by
PARTIAL YIELDING IN CYLINDERS 471

V (32-59)
V3 V «.

where r is to be computed for r = a from Eq. (32-58).


It is noted that the first terms in each of the three expressions Eq. (32-56) of the
stresses or,<rt,<r, represent three equidistant logarithmic curves belonging to the solu
tion of the flow problem for the case p — t, — 0, which was established in Chap. 31.
The second terms on the right sides of Eqs. (32-56) therefore represent "correction
terms" in first approximation, which must be added to the old solution to account
for the fact that Poisson's ratio v within the plastic zone o < r < c increases gradually
from its elastic value v' = 0.3 to its asymptotic value >>= H valid when the perma
nent strains became predominant. The correction terms in the expressions for a, and
a, in an example reproduced in Fig. 32-6 assuming that c = 4a turned out to be prac
tically insignificant. Only the curve representing the axial stress a, deviates from the
other a, curve, as can be judged by comparing the curve drawn by a thick line with
the dashed line for a, belonging to the asymptotic solution.
CHAPTER 33

THEORY OF THE PLASTIC DEFORMATION OF FLAT RINGS


OR DISKS

As mentioned in Chap. 31 a second group of rotationally symmetric


cases of the permanent deformations can be characterized as plastic
states of plane stress when the normal stress ot in the direction parallel
to the z axis can be taken equal to zero. This is the case in flat rings or
disks, the thickness of which is a small fraction of the inner or outer diam
eter. The components of stress o> and <rt in the radial and tangential
directions, being principal stresses, are functions of the distance r of a
point from the center of the ring.
33-1. Flat Ring of Ideally Plastic Material. Assuming small strains
and ideally plastic material, we have for the two unknown components of
stress <rr and <rt two conditions to satisfy:

a? — <rtffr + <7r2
= ffo* = const (33-1)
and

r
a7
= "' ~ °T ■ (33"2)

The first results from Eq. (30-15) by assuming in it a, — 0, and the


second is the condition of equilibrium. We note that a state of plane
stress in a disk may be determined from these two equations without
making it necessary to consider the permanent strains. Equation (33-1)
represents a plasticity ellipse in the variables ar and <rt having y/2 <r0 and
vHiro as major and minor semiaxis and whose principal axes bisect
the right angle of the o> and er( axes. This suggests again the use of two
and of an auxiliary variable angle 0:
<r'

variable stresses <r and

= £i+^= V2<r„sin0,
\/2
— (33-3)
_
\i(

C( 0>

(To cos
8

V2
by means of which o> and a, are expressed as follows:

(-0

a'

<r 2<Tn
r- sin
V3
(33-4)
2(T„
"*~

m
i)

V3"
\

472
PLASTIC RINGS AND DISKS 473

The auxiliary angle 6 may conveniently be used as the dependent variable

defined in Eqs. (33-3) satisfy the

<r'
in the ring problem. Since a and
condition of plasticity Eq. (33-1) which in the variables a and

a'
is
expressed by
<r2 + = 2<r„J W (33-5)
after substituting a, and the value of the difference of the stresses <rt ~ <r>
computed from Eqs. (33-4) in Eq. (33-2) the latter reduced to a differ

is
ential equation for the variables and 6:

r
_
d («-;)
.
rTsin cos . (33-6)

6
dr
The integral of (33-6)
is

-i = c_v/3* cos
(33-7)

6
where an integration constant. As in the problem of the thick-walled
is
c

long cylinder (Chap. 30) a differential


equation of first order obtained for
is

whose integral only one boundary con


dition can be prescribed.
Considering, for example, ring
p a

loaded by an internal pressure along


the surface = and having a stress-
a
r

free outer
surface = denoting
6,
r

by the value of Avhen = we


6b

6
r

have for the free edge the boundary


6-
condition: Flo, 33-1. Ellipse of plasticity.
= b, = 6b, er = (33-8)
0
r

and according to the first of Eqs. (33-4) we have to choose = x/6 or


6b 8b

ir + (71-/6). Point (Fig. 33-1) corresponds to the value of =


t/6 in
B

the ellipse of plasticity Eq. (33-1) or (33-5). If < x/6, ar will become
6

negative and the radial stress will be a compression. On the contrary,


if

x/6 < < <r> > and will be tensile stress. By keeping
a
0

71-/2,
6

we shall obtain a solution for a ring under internal pressure. The case
x/6 < < x/2 will correspond to a ring stressed by tensile forces in its
6

plane.1 Assuming =
x/6 the integration constant in Eq. (33-7)
is
9b

determined by taking =
r

This has been shown in Fig. 33-1 by the full and the dotted arcs of the ellipse.
1

The arc BCD of the ellipse represents the states of stress in the case of
ring with
a

internal pressure and the arc AB in a ring stressed by external tension (infinite plate
subjected to tension).
474 THEORY OF FLOW AND FRACTURE OF SOLIDS

,. -VI I cos t ,.
V3 o2e -Va I
= o2e 6 7;
6
= -tt
2
6

so that

*- 2
e^(i-0cos,. (33.9)

This prescribes on the inner surface r = a of the ring, where 6 = 8a, the

6tt*30 (Rings with infernal pressure)

SO" 6, \30° 0° -130"


I
-1/0'

Fia. 33-2. Radial and tangential stresses a, and inner circle r = a of flat rings
<ri at in
plastic equilibrium subjected to internal pressure p for various ratios b/a of outer to
inner radius.

condition determining the value of 0O:

\/3
6_2

cos oa
= (33-10)
a2
2

The pressure required for stressing the ring plastically found from
is
p

the value of o> = — for = a, = 6a:


p

6
r

,\
(r

2c0 .
(33-11)

Thus the complete solution for this case has been obtained in terms
of

the parameter
6.

value of the ratio b/a easily computed from Eq. (33-10) corresponds
A

to each value of (in fact the ratios b2/a2 are represented as a function
of
0„

the angle through one-half of wave of the familiar curve of a damped


da

oscillation). Furthermore value of the radial stress o> = — and


a
p
a

tangential stress vt corresponds for = a to value of Ba. Thus b/a,


a
r
PLASTIC RINGS AND DISKS 475

a, = —
p, and at are given as functions of 0„. This has been shown in
Fig. 33-2, in which b/a, o> = — p, and <rt at the inner bore r = a of the
ring are represented by the corresponding curves. The pressure p for a
ring with a given ratio b/a is read below the value of this ratio.
From Eq. (33-10) a conclusion can be drawn which is illustrated also
through the shape of the curve representing the ratio b/a in Fig. 33-2.
There is a limiting size to a flat ring beyond which it is impossible to
obtain complete yielding in it through applying a pressure on its inner

-1,2-1,0-0,8-0,6-0,4-0,2 O 0£ 0,4 0,6 0,8 I ISO.

Fig. 33-3. Distribution of radial and tangential stresses a, and at in flat rings in plastic
equilibrium subjected to internal pressure p (a inner, b outer, r variable radius, at yield
stress in tension).

boundary. The ratio b/a has an analytic maximum. This is easily


computed from Eq. (33-10) by differentiation. It occurs at 6a = — t/3

and is equal to (6Vo*)„« = e »/V'3 or (b/a)^ = 2.963. Should a


ring have a ratio of outer to inner radius b/a which is larger than 2.963, the
outer portion of the ring must remain strained elastically and a case of
partial yielding is obtained.
A second difference between the behavior of plastic rings and of long
cylinders may be mentioned. The internal pressure p under which a long
cylinder with closed ends («* = 0) yields increases indefinitely with
increasing ratio b/a. In a flat ring subjected to a uniform radial pressure
on the inner surface r = a, on the other hand, the pressure p as well as
the tangential stresses can never exceed the maximum value

2<To
= 1.155<to
V3
(16 per cent above the yield stress o-0 in simple tension).
476 THEORY OF FLOW AND FRACTURE OF SOLIDS

The distribution of stress in plastic rings is shown in Fig. 33-3 in which


the radial and tangential stresses are plotted in function of the radial
distances r from the center of the ring. These two stress curves are the
same for all ratios b/a ending abruptly at b/a = 2.963. Figure 33-4

1 i I 1 1 f °\

Fig. 33-4. Stresses ar and <rt in thin disk having circular hole of radius a stretched by
uniform stress ffo on outer circumference r = « .

reproduces the distribution of stress in an infinite plate uniformly stressed


in its plane at r = oo by a tensile stress <r> = at = <ro in all directions and
having a circular hole. In this case 30 deg < 0 < 90 deg. See also
Sec. 33-3.

33-2. Permanent Expansion of a Hole in a Steel Plate. In industrial fluid con


tainers serving as heat exchangers and in the high-pressure boilers and the large
condensers of steam turbines, small tubes are fitted in large numbers to drums or

SECTION A-A
Fio. 33-6.
Fios. 33-5 and 33-6. Tube expander.

headplates by expanding slightly permanently the tube ends by means of the pressure
of small tapered rolls (Figs. 33-5 and 33-6). A cross section through a tube joint is
reproduced in Fig. 33-7. The radial enlargement of the tube ends depends on the
pressure difference which the joint must sustain without leaking. Until recently
empirical data were available only concerning the necessary enlargements of the con
tact radii. The theory of the ideally plastic substance can furnish engineers this
PLASTIC RINGS AND DISKS All
information,1 since the steel used in similar applications very nearly conforms with
the assumptions on which the former is based. An expanded tube joint maintains
its tightness against a certain pressure difference through the system of residual stresses
which remains around the contact area due to the small permanent distortion of the
steel plate. Assuming an infinite steel plate
of constant thickness having a circular hole
of radius a one part of the question consists
in the determination of the distribution of
stress around the hole during the permanent
enlargement of it and of the function connect
ing the pressure at the hole with the radial
enlargement. This is a case of partial yield
ing. Upon release of the radial pressure ex
STEEL
erted by the expanding tool a second part of PLATE
the analysis is the determination of the sys
tem of residual stresses remaining around the
joint and of the contact pressure. Only the
first part of this problem is briefly described.
Assuming a well-defined yield point and a con Flo. 33-7. Tube and steel plate.
stant yield stress a„ under a certain radial
pressure p there will develop a plastic zone in the steel plate a g r £ c which is sur
rounded by an elastic zone c £ r < », in which the stresses are equal to

ffo C
(33-12)
V'3 r

and in which the elastic radial displacement u is

(l + i.)<roc«
u — err = (33-13)
V3 Er

(E modulus of elasticity, v Poisson's ratio).


Theexpre&sions for the stresses a, and at and the function connecting the variable
radius r with the parameter 8 have already been established for the plastic zone
a < r < cm Eqs. (33-4) and (33-7) in Sec. 33-1. We may now interpret the integra
tion constant c in Eq. (33-7) as the variable radius c of the boundary circle between
the elastic and the plastic: zone. However, we must choose a new boundary value 8e
for the angle 8 when r becomes equal to c. Since for t = c, trt = —a, = <ro/\^3 from
Eqs. (33-4), we see that we must make 8C = 0, 8 < 0, and

-\/30 cos 8 , (33-14)

When 9=0, yielding starts in the plate at r = c. To 8 = 0


corresponds the point
C in the ellipse of plasticity (Fig. 33-1). As r varies between the values a g r g c
the angle 0 varies between the values 9„ g 8 g 0 and the condition
-y/iOa
(33-15)

1 Trans. ASME, Davenport, Iowa, meeting, April, 1943. Also Goodier, J. N.,
G. J. Schoessow, Power Division, ASME 1942 meeting. Grimisox, E. D., and G. H.
Lee, ASME 1942 meeting.
478 THEORY OF FLOW AND FRACTURE OF SOLIDS

determines the angle 8„. A value 8a and a pressure p correspond to each value of c at
which yielding has penetrated in the plate to the radius c. The pressure p can never
become larger than 2o0/\/3- The following table gives the values of the angle 8C and
the pressure p at the bore r = a:

8a, deg Pressure p Interval for 8, deg

0 0.500 X 2cro/\/3 0
-15 0.707 X WV3 -15 < e <0
-30 0.866 X aro/v/3 -30 < e <0
-45 0.966 X WV3 -45 < e < 0
-60 1 X WV3 -60 < 8 <0

There is no need to carry the radial enlargement of a hole in the steel plate beyond
the maximum value of the pressure p = 1.155<ro because only an excessive deformation
in the joint would result without gaining in the amount of the residual contact pres
sure (assuming that in a very ductile steel the strain hardening has not yet developed
appreciably within the range of these strains). A few curves in Fig. 33-8 illustrate
the distribution of the radial and tangential stresses o> and m in the steel plate around
the hole under increasing values of the radial pressure p in the plastic and elastic
region of the plate. ,
In order to determine the pressure p as a function of the radial enlargement u« of
the hole we shall assume that, in the plastic region a g r g c, stress-strain relations
for an incompressible material hold of the form

(33-16)

(assuming that a, = 0 where «r,«( denote the small permanent strains in the radial
<t>

and tangential directions, u the radial displacement, and the flow function, neglect
ing in this approximate consideration to separate the strains in their elastic and perma
nent parts.1 Since the components of stress <r, and at have already been determined
in Sec. 33-1 [Eqs. (33-4) and (33-7)], the condition of compatibility of the strains

cU,
(33-17)
<t>

makes possible to compute an expression for the function after substituting


it

in

Eq. (33-17) the values of «r and «< from Eqs. (33-16). This gives

— —
*^2<7' **) = 3*(°v "') (33-18)
r

dr

Instead of using relations of the form of Eqs. (33-16) the increments of the strains
1

dt,,dti should have been introduced and their integrals computed as the pressure
p

increases with the advancing plastic boundary = c. The determination of the


r

radial enlargement of the hole u„ in the text above represents an approximate way of
estimating «„.
PLASTIC RINGS AND DISKS 479

0.75

0.25

-0.25

-0.5

-0.75

-1.25
Fits. 33-8. Stress distribution in steel plate having a hole after partial yielding through
radial expansion.

After making use of the auxiliary variable 0 introduced in Eqs. (33-3) and (33-4)
and of Eq. (33-6) this last equation reduces to

(33-19)
<f>

= —
<//

where cos (»/6)J. Therefore,


[9

and
-V3»
COS
Coe
- (»/6)] (33-20)
[0

The strains c, and n are

e~^e cos
- (»/6)]
+ (ir/6)]
[0

— Cairo
cos (33-21)
[0

-y/3»
ci = Caaae
480 THEORY OF FLOW AND FRACTURE OF SOLIDS

The integration constant Co is determined by equating the value of the strains in


the boundary between the elastic and plastic zone r = c, 8 = 0 computed on the
elastic side from Eqs. (33-12) and on the plastic side from Eqs. (33-21) (taken for
e = 0):
(1 + »Vo
c0 =
1 +*
y/ZE \ 3 E
(33-22)

After dividing Eq. (33-14) by Eq. (33-15) we see that

COS ffa
(33-23)
cos 8

The radial displacement u is given by the second of Eqs. (33-21) using the preceding
expressions
u = e,r = rwe-Vs* . (33-24)

The radial enlargement u. of the hole r = a is therefore found equal to

= (1 + "Vo"
u e-V3B. (33-25)
V3 E
when the pressure p has the value

V3
Figure 33-9 shows the curve representing the function p = /(»„) according to
Eqs. (33-25) and (33-26) and how the pressure p increases with the radial enlarge
ments M„ of the hole. The first branch OA corresponds to a purely elastic expansion;

J 4
RADIAL EXPANSION
Fio. 33-9. Plastic expansion of a hole.

the plastic branch of the curve begins at point A and ends abruptly with a horizontal
tangent at point B when p reaches its maximum value 2<T„/v 3 = 1.155<r„ and the
radial displacement u„ has become equal to 6.12u„. na is the radial displacement at

r = a when the plastic limit is just reached in the surface r = a of the hole (point A in
Fig. 33-9), which is equal to u„ — (1 + >0<roa/\/3 E.1
1 The second curve in Fig. 33-9 shows for r = a the permanent increase in thickness
of the plate as a function of M„. The ordinates of this curve are proportional to the
strain e,. It is instructive perhaps to note that the ordinates of this curve increase
very rapidly with the expansion ua, thus limiting also for this reason the pressures p
PLASTIC RINGS AND DISKS 481

Figure 33-10 shows the distribution of the residual stresses in a tube joint after the
steel plate has partly yielded to a radius r = c and upon release of the expanding
pressure. The ordinate between the points B and C denoted by p indicates the residual
normal pressure remaining in the contact surface between the tube and the steel ■plate. The
curves denoted by ar* and a* represent the
distribution of the stresses at the instant
when yielding was interrupted. The curves
denoted o> and it, represent the stresses in
an elastically strained tube joint under a
pressure equal to the pressure at which
yielding was interrupted, and the differences
of the ordinates of both curves (the shaded
ordinates) visualize the distribution of the
residual stresses.
33-3. Infinite Plastic Disk in Tension
Having a Circular Hole. As already men
tioned, a further simple case included in Eqs.
(33-4) and (33-7) is the plastic distribution
of stress in an infinite disk stressed uni
formly in its plane in all directions by ten
sion and having a circular hole. This case
is obtained by choosing for the variable 6
the interval 30 deg < 9 < 90 deg. The
distribution of stress in the disk is repre
sented in Fig. 33-4. It is well known that
in an elastic disk stretched in its plane by a
stress a> = a, = a the stress a, along the cir
cumference of a hole is equal to 2a. This is
usually expressed by stating that a circular
Fio. 33-10. Residual stress distri-
hole in an elastic plate stretched uniformly bution in a tube joint after partial
in all directions produces a concentration of yielding of the steel plate.
the stresses. The factor of stress concen
tration denned by the ratio A; = a, /<r for an elastic disk is equal to 2. This ratio
k is thus reduced from 2 to 1 when the disk yields completely since the tangential
stress a i at the surface of a hole is just equal to the yield stress ac in pure tension and to
the stresses at large distances from the hole in an ideally plastic material.1
1 Equivalent simple formulas for the stresses in an infinite disk could have been
obtained by "linearizing" the condition of plasticity Eq. (33-1) by substituting for
the arc AB of the ellipse of plasticity (Fig. 33-1) a horizontal straight line. If the
secant AB itself is substituted, this is equivalent to stating that the solution is derived
"
from the "maximum shearing stress theory of flow. (In a disk stretched by uniform
tension or and a% are everywhere positive and a, is the largest principal stress, the
smallest being equal to zero; therefore rmu = <xi/2 = c/2 = const.)
CHAPTER 34

DISTRIBUTION OF STRESS IN ROTATING CYLINDERS


AND DISKS

The theory of the distribution of stress in rotating cylinders or in


revolving disks after the yield point has been reached is similar to that
of the stress distribution in thick-walled tubes or flat rings treated in the
preceding chapters. That such questions have practical importance is
indicated by the fact that engineers have long found it necessary to specify
a high degree of ductility in the materials used for machine elements such
as the rapidly revolving disks or heavy spindles of steam turbines or the
heavy cylindrical rotors of large turbogenerators, which are mainly
stressed by the centrifugal forces. In overspeed tests of such highly
stressed cylinders or disks the yield point of the material may be reached
or passed in certain portions of the disks. As A. Stodola mentions,1
attempts have been made to improve the stress distribution in rotating
disks with a central hole by running them during manufacture at a speed
such that the inner part of the disk is permanently stretched. This
question has also been considered by H. Hencky, F. Laszl6, and others.8
Therefore it will be perhaps useful to work out in what follows some of the
simplest cases of plastic flow in rotating cylinders or disks.
34-1. Rotating Cylinder under Plane Strain. In this case the unit
elongation in the direction parallel to the axis of the cylinder e, = const
and according to Sec. 30-1, Eq. (30-2), the radial displacement p is
given by

P=-(i + J- (34-D

Now in a solid cylinder, which case will be treated first, the constant c\
must be chosen equal to zero: d = 0. In a solid cylinder the radial
displacement p and the tangential and radial strains t < and e, must be

p - - |r , (34-2) «, - '- = -§ , (34-3)

1 Loewenstein, L. C, "Steam
and Gas Turbines," Eng. ed., vol. 2, p. 1080.
McGraw-Hill Book Company, Inc., New York.
* See a paper by L. H. Donnell and the author in
Tran*. ASME, 1928, on Rotating
Discs, where a bibliography on this subject can be found.
482
REVOLVING (OR ROTATING) CYLINDERS AND DISKS 483

We see that in a solid rotating cylinder in the plastic state, the deforma
tion in each element is practically the same. It consists evidently of a
unit elongation t, = const parallel to the axis of the cylinder which is a
uniform contraction and negative, accompanied by a lateral elongation
of half this amount and equal in all directions perpendicular to the axis
of rotation which will therefore be a uniform expansion in the radial
direction.
According to the third rule of plastic flow, we note that since er = «,,
also at every point the radial and tangential stresses must be equal,

a, = a, (34-5)

as otherwise the figure ofMohr's principal stress circles would not remain
geometrically similar to that of the principal strain circles (see Chap. 16).
The condition of plasticity

(ar
- <r,)2 + (a, -a.Y+ (a, - <rr)2
= W (34-6)

becomes therefore simply

a, — a, = +<r0 = const , (34-7)

where is the yield stress in pure tension.


ao In the condition of equi
librium an additional term must be introduced to take into account the
action of the centrifugal forces. Let to be the angular velocity of the
cylinder, X the specific weight of the material (weight in pounds per cubic
inch) g the acceleration of gravity, and a the radius of the cylinder. Then

dr r g
v '

According to Eq. (34-5), the first term on the right side of this equation
is now zero and the equation of equilibrium takes the simple form

da, Xco

whence by integration
X«S ~
Cr = rS) (34-10)
W (fl2

using here the condition that for r = a, ar = 0. The tangential stress


at is according to Eq. (34-5) equal to o>.
According to Eq. (34-7) two distinct cases may be obtained, depending
on whether the upper or the lower sign is taken. Taking the upper sign,
the axial stress is given by

<r, = *,-<ro. (34-11)


484 THEORY OF FLOW AND FRACTURE OF SOLIDS

In a freely rotating cylinder, which can contract freely in an axial direc


tion the resultant of the axial stresses a,

2wf\,rdr = 0
(34-12)

should become zero. From this condition the peripheral speed u = aw

is found under which the entire cylinder will yield

u = aw = 2
\

(34-13)
^

Now the peripheral speed u0 of a free ring of small cross section rotating
with a uniform angular velocity and stressed to a stress <r0 is given by the
expression

wo
= (34-14)

and this is the speed under which in a free ring the yield stress <ro will
just be reached. Hence
u = 2w0 . (34-15)

The distribution of stress under which the whole rotating cylinder


becomes plastic is finally given by the set of formulas

Radial and tangential stress: = — 2<T0 —


<rr <rt (34-16)
^1

Axial stress: a, = <r„


-
(34-17)
^1—2 ^\
The central part of the cylinder receives tensile and the region near the
periphery compressive stresses in the axial direction, but the action of the
former balances that of the latter, so that no resultant axial force is acting.
34-2. Comparison with Elastic Rotating Cylinder. In free solid
cylinders of an elastic material of radius a rotating with an angular
velocity to or with a peripheral velocity u' = aw the stresses are, accord
ing to the theory of elasticity, given by

— (3
- 2v)<r (. r*\

<ri
-
sir
_ 2„ - (1 + 2y)
£
. (34-18)
[3 |
a* ~
4(T
REVOLVING (OR ROTATING) CYLINDERS AND DISKS 485

where v is Poisson's ratio and a stands for the constant

..^-V g
tfj.
wo2
(34.19)

For steel, for example, if v = %, the stresses at the center of the cylinder
are
r = 0 , trr = ot = 0.437(7 , cr, = 0.125* (34-20)

and at the periphery of the cylinder

r = a, <yT = 0 , at = 0.125a , <r, = -0.125a , (34-21)

and first yield will occur at the center r = 0 of a steel cylinder when at
this point

»i-».- ?/i~ —4y)r - »• i * = 3.20*0 . (34-22)


o\l v)

Hence a solid rotating steel cylinder of radius a will start to yield at the
center if the peripheral velocity u' is equal to

W - *P =«o 2u0 J2„(1


~ y)
= 1.789uo . (34-23)

Above we found that the whole cylinder becomes plastic at a speed


u = 2u0. Comparing u with u' it is seen that from the moment at which
the first yield occurs, an increase of speed of about 12 per cent will bring
the entire cylinder into the -plastic state. The plastic region, therefore,
spreads comparatively rapidly with changing speed, and an increase of
only 12 per cent suffices to obtain plastic deformation in the entire
cylinder, provided that its material follows the law of deformation which
was assumed, i.e., that the material yields under a constant stress <ro in
pure tension.
The stress distributions in these two cases — for first yielding and for
complete yielding — are shown in the Figs. 34-1 and 34-3.
We may mention, that the case of a hollow rotating plastic cylinder
can also be worked out completely without difficulty, but the expressions
for the stresses become slightly more complicated.
34-3. Rotating Disk. In a rotating disk having a uniform thickness,
which is a comparatively small fraction of its diameter, the approximate
value of the stresses can very easily be found in the cases of both partial
and complete yielding. As in this case one principal stress, namely, a,
vanishes throughout the disk, the two other principal stresses in the
plastic region according to Sec. 33-1 must become the rectangular
coordinates of an ellipse. Now in a disk stretched by centrifugal forces

'
486 THEORY OF FLOW AND FRACTURE OF SOLIDS

both stresses a, and <r, will be tensile stresses and as the tangential
stress always remains the larger, we may replace the arc of the ellipse
expressing the condition of plasticity used before (see Eq. (33-5), page 43,
and the arc of the ellipse in Fig. 33-1 between A and B) simply by a
horizontal line. The condition of plasticity will therefore be expressed
with sufficient approximation by

tTl
= c = const . (34-24)

As the value of the constant c we may take either the average value of

TANG.

Fig. 34-1. Fig. 34-3.

Fig. 34-2. Fiu. 34-4.

0 —"1 PLASTIC
Figs. 34-1 and 34-2. Elastic equilibrium. Stresses in rotating cylinder and disk just
when first yielding starts.
Figs. 34-3 and 34-4. Plastic equilibrium. Stresses in rotating cylinder and disk when
entirely in plastic state.

the yield stressot> in pure tension (which is the ordinate of point B in


Fig. 33-1) and of the maximum possible ordinate of the ellipse (which
is = 2ff0 \/3 = 1.155<7o) by taking

at = c = ! -^=Vo = 1.077<ro (34-25;


^2 V3/
or we may use the theory of maximum shear, according to which <r, also
should be chosen equal to the constant:

ci = co = const . (34-26)
REVOLVING (OR ROTATING) CYLINDERS AND DISKS 487

(We see that the above approximate value of the constant c differs by
8 per cent from the value obtained from using the theory of maximum
shear.)
a. Solid Disk. In the case of a partly plastic solid rotating disk of
radius o we have an outer elastic portion and an inner plastic region.
As long as the stresses are below the yield point they are given by the
expressions derived from the theory of elasticity, namely,

(3 + v)c
8 0-5).
*, -| + v - + 3.)
(1
(34-27)

[3 £]
a, = 0 , (34-28)

where again v denotes Poisson's ratio and a stands for the constant

j-*-*'.^, 9 Wo2
(34-29)

(u' peripheral speed of disk, m0 speed of free rotating ring, which yields
under stress a0).
The disk yields first at its center r = a, at a peripheral speed u' found
by equating at for r = 0 to <r0.
This gives

a = „
"
and v! = m0 */— = "o *L ,
= 1.55u0 . (34-30)

If u' > 1.55u0 there will be an elastic portion extending from a radius
r = c to r = a in which the stresses will be equal to

, (3 + »V r» c,
^ +
g^ a* ci-7*'
(34-31)
(3, + IV r» c,
zi + ci + r?

ff« 5
8 a? Cl+7*
and a plastic portion extending from r = 0 to r = c, in which according
to the theory of maximum shear

~ (rr2 \
ffr ~ *° '
3Tl (34-32)
Cj = ffo .
J
/

The three unknown constants a, ch ct can be found from the conditions


that for r = a, <rr = 0 and for r = c, <rr = o>', <r( = <r,'. Using the theory
of maximum shear (<r, = (To = const), the peripheral speed u can be
488 THEORY OF FLOW AND FRACTURE OF SOLIDS

computed explicitly at which the plastic portion will extend to a given


radius r = c

-
24a4
M2 = U°2 '
(34_33)
3(3 + y)a* 2(1 + 3i)oV + (1 + 3«)c<

From this if c = a we see that the entire disk will yield if the peripheral
speed will be equal to
u = a/3 Mo
= 1.73u0. (34-34)
Comparing this last value with u' = 1.55«o which was the speed for first
yield, we see that an increase of about 12 per cent in the speed is sufficient
to spread the plastic state through the whole disk. Figures 34-2 and 34-4
show the state of stress in the rotating disk at the moment of first yielding
and for complete yielding.
b. Rotating Disk urith Hole. In this case an additional term has to be
introduced in the expression for the radial stress o> within the plastic
region a < r < c:

Using again the theory of maximum shear the peripheral speed under
which the disk will yield to a radius c is found

I2[2b'c
- ab*(b* + e2)]
^
U* = M°2
3(3 + v)b"c
- (1 + 3i0(2fc*
- c*)c3
- 4a3(62 + c2)
'
t34"30)

Here a and b designate the inner and outer radius of the disk and «o the
speed in a free ring of small cross section for yielding. (u02
= <rog/\, o-o
is the yield stress in pure tension, g the gravity acceleration, X the specific
weight of the material of the disk, v is Poisson's ratio.) According to the
last formula the first yielding starts at the inner surface r = a at a
peripheral speed u' = ub given by the value of u i or c = a, which is

3 + v + (1
- y)(a2/62) (Mr*<)

and the plastic region will spread completely to the outer surface at a
peripheral speed u given by the value of u for c = b, which is

We may verify this last result by computing the value for u for the case
that a tends to become equal to b. In this case the disk becomes a thin
ring with small thickness in a radial direction. Formula (34-38) gives for
a/b = 1,

u = Mo VH = Mo (34-39)
REVOLVING (OR ROTATING) CYLINDERS AND DISKS 489

in agreement with the speed u0 for a free rotating ring of small cross sec
tion and just at the yield point. Figure 34-6 shows an example of a stress
distribution in a rotating disk with a hole and in a partly plastic state.1
The preceding developments have to be generalized for a metal exhibiting pro
nounced strain hardening by introducing a strain hardening function t0 = f(y<>) in
the manner discussed in Chap. 31 if the stresses in rotating cylinders or disks beyond
the elastic range of the strains have to be
considered. A related problem has been INNER HALF WHOLE DISK
PLASTIC JUST PLASTIC
discussed by F. K. G. Odquist,* who
treated the case of the slow creep at ele i.fA
vated temperature in a rotating disk of
variable thickness having approximately
the profile of a body of revolution with
two flat conical surfaces. He assumed a
relation in the form t„ = f(ya) or that the
octahedral shearing stress t„ is a function
of the rate of the octahedral unit shear
7„, for which he used a power function
t = cy0m. His theory, according to what
has been stated (Sec. 28-3), appears appli ELASTIC ELASTIC PLASTIC

cable to a rotating metal disk in the plas Fio. 34-5. Fio. 34-6. Fig. 34-7.
Figs. 34-5, 34-6, and 34-7. Rotating disk
tic state whose material would yield
with hole. Radial and tangential stress
according to a power function t„ = f(ya)
o> and <T|. <ro yield stress for tension.
as strain hardening function.
The case of the yielding of a rotating disk of an ideally plastic material having an
arbitrarily given profile has been treated by A. S. Thompson,3 who assumed a gradient
of the temperature in the radial direction in the disk and a variable yield stress and
modulus of elasticity, both of which he assumed dependent on the distribution of the
temperature in the disk.
Some tests made with steam turbine disk wheels by speeding them up until they
burst in pieces were described by E. L. Robinson.* The permanent distortion of
disks of equal thickness revolving at high speeds in an evacuated chamber (for reduc
ing the friction created by the air) was studied by C. W. MacGregor and W. D.
Tierney • in connection with the strength of welds in steels and for the purpose of investi
gating the fracture of steels under a state of equal biaxial tensile stresses occurring in
the center portion of the rotating disks. In a solid disk of 13 in. radius and 0.5 in.
thickness of hot rolled 0.23 % C steel they observed that the fracture started as a
radial crack and a shear fracture in the center portion of the disk but changed to the
herringbone cleavage fracture in the outer portions of the disk after the disk burst at
11,800 rpm, as was observed in the tests made with tubular specimens by E. A. Davis
described on page 190.
1 In
the paper mentioned above (see footnote 2, p. 482) L. H. Donnell has worked
out several cases numerically under the assumption used above and also by using the
more general condition of plasticity.
5 Proc. Royal Swedish Institute
for Engineering Research, No. 141, 1936, Stockholm,
31 pp.
' J. Applied Mechanics, ASME meeting, 1945.
4 Trans. ASME, 1943 meeting.
6 Welding Research SuppL, vol. 8, No. 6, pp. 303-309, June, 1948.
CHAPTER 35

THE PROBLEM OF PLASTIC TORSION. EXPERIMENTAL


REPRESENTATION OF STRESS DISTRIBUTION
Yet who does not prize these moments that reveal to us the poetry
of existence.
Eddington, Arthur Stanley, "Science and the Unseen World,"
Swarthmore lecture, 1929, p. 46, The Macmillan Company, New York.

The transition from a state of equilibrium with purely elastic deforma


tions to one which may be postulated in an ideally plastic material can be
studied generally in a cylindrical or prismatic bar of constant cross sec
tion twisted about its axis. Although the methods presented in this
chapter describe the distribution of stresses in twisted bars strained above
the yield point by utilizing visual rather than analytical means, they offer
the basic equations for evaluation of the distributions of stress. Several

I"
SHEAR
STRESS

UNIT SHEAR

Fio. 36-1. Via. 35-2.


Fio. 35-1. Idealized stress-strain diagram.
Fig. 35-2. Section through twisted bar.

cases have been analytically and numerically computed in recent invest i-


gations. As in previous chapters for metals having a sharply denned
yield point, an idealized stress-strain diagram may be introduced for the
beginning of plastic flow consisting of two straight lines as shown in Fig.
35-1, in which the steeply inclined straight line refers to the elastic and
the horizontal line to the plastic range of strains. The states of stress in
the small elements of a prismatic bar subjected to pure torsion are taken
as states of simple shear. As long as the shearing strains remain quite
small, the assumption will be made that no components of normal stress
act in the sections normal or parallel to the bar axis and that the resultant
of the two components of the shearing stress t„ and tvz, which are the only
components acting in the cross sections perpendicular to the axis of the
490
PLASTIC TORSION 491

bar everywhere in the section where the yield point has been reached, has
a constant value A; (the constant k is equal to <r0/2 if the maximum shear
ing stress theory is assumed and equal to <r0/\/3 if an ideally plastic
substance is postulated yielding under a constant octahedral shearing
stress, where a0 is the yield point for tension). In the cross sections of
the twisted bar in which the metal yields partially there will be certain
regions which are strained only
elastically and plastic regions in
which this stress has reached the
limiting value k.
In order to analyze the distribu
tion of stress in a twisted bar after
the yield point has been reached, it
is first necessary to recapitulate how
this distribution is determined in an
elastic bar.
35-1. Elastic Torsion. Prandtl's
Soap-film Analogy. The Elastic
Stress Function for Torsion. The
distribution of shearing stresses in
an elastically twisted bar may best
be visualized by using Prandtl's
membrane or soap-film analogy. In
order to find the resulting shearing
stress which exists at a given point
P of a cross section of the bar Fig. 35-3. Ludwig Prandtl (born 1875).
having the rectangular coordinates
x,y,z referred to a system whose origin has been chosen at a point of the
axis around which the bar is twisted and whose z axis coincides with the
bar axis (the point 0 in Fig. 35-2 represents its intersection with the
plane of the figure), let the resulting shearing stress r at point P be
resolved into the perpendicular components rz and t„ in the directions of
the x and y axes.1
The equilibrium of stresses acting on an element, dx, dy, dz is expressed by the
equation
.I, A,
dr» . dry
= 0 . (35-1)
dx dy
1 Since these two components of the shearing stress are the only ones acting in the

plane of the cross section and on the faces of an element dx,dy,dz the symbols t„tv are
taken to represent rM and t„„ respectively. For the theory of elastic torsion rf.
Love, A. E. H., "Mathematical Theory of Elasticity," 4th ed., Chap. 14, Cambridge
University Press, Ixmdon, 1927; or Timoshenko, S., "Theory of Elasticity," Chap. 9,
p. 228, McGraw-Hill Book Company, Inc., New York, 1934.
492 THEORY OF FLOW AND FRACTURE OF SOLIDS

Let f ,q, f
be the small displacements of the point x,y,z relative to its initial posi
tion and parallel to the axes x,y,z. Under twist the cross section of the bar z constant
rotates about the z axis. This is expressed by the equations

{ 6yz , v = Bxz , r = H(x,y) . (35-2)

These displacements show that the projection of the point x,y,z, on the x,y plane
describes a small arc of length 6r (where r* = x* + yl) with respect to the z axis and
that the originally plane cross sections of the bar are, in general, distorted into sur
faces given by the function <t>(x,y). The constant 6 in Eq. (35-2) is the angular twist
of the bar per unit of length.
We now express the unit shears y,, and yu. by using Eqs. (11-13) as follows:
-
J. ^ " \
y" = 9<
ol
+ ax
«
9
U
(d* ~
y) '
(35-3)

The shearing stresses t, and t„ are given by Hooke's law:

(35-4)

In this G represents the modulus of rigidity.


From these equations it follows that at each point x,y of the cross section, the follow
ing equation must hold :

_ ^£ - -2G9 = const . (35-5)

In order to satisfy Eq. (35-1) the shearing stresses are taken equal to the derivatives
of a certain function F(x,y) :

T' = Jj • T' =
~
ai

(3M)

The function F(x,y) is called tfie elastic stress function of the cross section.
From this, using Eqs. (35-5) and (35-6), the following partial differential equation,
denning the function F , is obtained:

^ + ^ - AF 2G8 = const . (35-7)


dx* dy*

In order to obtain the boundary conditions, i.e., the conditions which the function F
must satisfy at the edges of the cross section, it should be considered that the resulting
shearing stress t at each point of the boundary of the cross section shall have no com
ponents perpendicular to the edge. Let y = f(x) be the boundary curve of the cross
section. Therefore, along the boundary curve y = f(x) the following equation must
hold:

^ = ^ , (35*)
T, dx
PLASTIC TORSION 493

by which the fact that the shearing stress t is tangent to the curve y = f(x) is expressed.
Substituting Eq. (35-6) in Eq. (35-8), there results

dF
— dF

-tv dx + t, dy = dx + dy = 0 . (35-8o)

This means that along the boundary curve y = f(x) of the cross section the ordinates
of the functionF must be taken as constant.
The twisting moment M is expressed by the double integral

M =
JJ (tvx
— r,y) dxdy = — jJ x
+j^y^
dxdy

and if the constant value of the stress function F on the boundary curve y = f(x) of
the cross section is taken equal to zero, by partial integration:

J j ~xdx f dy(xF J
dy = — F - — dxdy,

dI)

F
JJ
similarly,

dx = - dxdy

F
Jj
j

^ydy
J

the moment found equal to


is

M - Fdxdy
Jj

. (35-9)
2

These equations contain the essentials of the Prandtl membrane or soap-film


analogy which as follows: thin membrane stressed uniformly in its plane soap
A

(a
is

film) to be thought of as attached to the boundary curve of the cross section of


is

a
twisted bar. This membrane loaded by constant external pressure. It can easily
is

be shown that the deflection of such a membrane would satisfy differential equation
a

of the form Au = const, while the deflection u has a constant value along the edge.
The curved surface into which the membrane distorted by the lateral pressure and
is

which may be called the soap-film surface of the cross section thus satisfies essentially
the same conditions as the function F(x,y) or the elastic stress function of the cross
section. Along the contour lines of this surface u = const or = const. At every
F
is

point along a contour line according to Eq. (35-8a)


:

§M _ _ f.*? = L„ .
t, (35-10)
dx dx dy

This means that the resultant shearing stress r tangent to the contour lines
is

F{x,y) = const
.

Furthermore, we have

'.--(©• +©'.
This the square of the largest slope of the surface (the gradient of the surface)
is

F(x,y). The resultant shearing stress at any given point x,y of the cross section
is
r

therefore equal to the greatest slope of the stress surface F(x,y) at this point.

Briefly recapitulating, the preceding membrane analogy of elastic


torsion may be expressed as follows: thin membrane thought of as
A

is
494 THEORY OF FLOW AND FRACTURE OF SOLIDS

fastened along the boundary curve of the cross section and loaded by a
uniform surface pressure. The curved surface into which the membrane
is distorted is called "the soap-film surface" or "the elastic stress func
tion" F(x,y) of the cross section. The contour lines of the elastic stress
surface F(x,y) represent the stress lines of the cross section of an elastically
twisted bar. At each point the tangent to the contour line gives the
direction of the resulting shearing stress t, while this stress itself is pro
portional to the greatest slope of the stress surface F(x,y). The twisting
moment M acting on the bar is equal to twice the volume enclosed by
the stress surface F(x,y).
35-2. The Plastic Stress Function of Torsion. The Sand-heap Anal
ogy.1 If
an iron bar is severely twisted, certain parts of the bar will
undergo plastic deformation. At a certain point of the cross section at
which the shearing stress t has reached the yield point, both shear-stress
components tx and t„ must satisfy the condition of plasticity:

T,* -I- V = fc2 = const . (35-12)

The last equation states that of all shearing stresses occurring at various
cross sections of the bar, the largest, which in this case is t, has according
to the assumptions made (see Fig. 35-1) at the yield point a constant
value These components t, and t„ must further satisfy the condition
fc.

^
of equilibrium
=
+

Tx

Ty
This equation again satisfied we put
is

if

The function F(x,y) determined by these relations may again be repre


sented as surface over the cross section. F(x,y) may be called the plastic
a

stress function the cross section. In those parts of the cross section
of

in

which plastic flow occurs, the plastic stress function must satisfy,
F

according to Eqs. (35-12) and (35-14), the following equation:


(f)

=fc2.
+(|D (35-15)

The differential expression on the left side of this equation the square
is

of the absolute value of the gradient "grad F" (of the maximum slope

lZ. angew.Math. Mechanik, vol. no. p. 442, 1923. Cf. also Plastic Torsion,
3,

6,

an Experimental Determination of the Stress Distribution in Bar Which Has


a

Twisted to the Limit of Plasticity, Proc. ASME Mech. Division, 1931.


PLASTIC TORSION 495

of the surface Everywhere in the cross section where flow occurs the
F).
following equation must hold true:

|grad F\ = k = const .

By means of this property and the further condition


rip1 rlF
-Tydx + T,dy = + ~dy = 0, (35-16)
^dx
according to which the shearing stress t at each point along the edge of a
plastically distorted part of the cross section is directed tangential to the

Flo. 36-6. Fig. 35-7.


Flos. 35-4 to 35-7. Examples of plastic stress function for torsion represented by wooden
models for various cross sections.

edge fix),1 or the condition that along the edge F = const, the
y =
plastic stress function F of the cross section is determined. Since an
additive constant in F does not affect the value of the stresses (the shear
ing stresses are the derivatives of F) along the edge, F may be taken equal
to zero.
1
Cf. Eq. (35-8a).
496 THEORY OF FLOW AND FRACTURE OF SOLIDS
*

From the above-mentioned properties of F it will be seen that the plastic


stress function is a surface of constant maximum slope which one may
construct over the edge of the cross section. If the contour of the cross
section is thought of as cut out of a piece of stiff paper and covered with
sand while lying horizontally, there results a heap
whose natural slope gives a picture of the surface F.
Its form is independent of the amount of twist (the
angular twist 0). The shape of the plastic stress func
tion F(x,y) is represented by some wooden models in
Figs. 35-4 to 35-7.
We have now to consider how the plastic areas Ai,
Uo°Uyt,d3£ A* A,, ... of the cross section (Fig. 35-8) may be
tic (A j, Ai) regions differentiated from the elastic area Ai at any given
in cross
secUor^of vame Qf twisting moment. We designate with the sub-
above the plastic script (1) all values which refer to the elastically dis-
lltnit-
torted parts of the cross section. With respect to the
elastic stress function F\, we know that its ordinates taken over the elastic
part A\ of the cross section satisfy the following differential equation:1

-a^+W'^'-208- (35-17)

Along the of the cross section the condition F\ = 0 is satisfied.


edges
(G represents the modulus of elasticity in shear, 9 the unit angular twist
of the bar.) We assume for a moment that the values of Fx are known
along the boundary curves of the plastically distorted areas. Then the
surface Fx is determined and it is possible to draw the contour lines
F\ = const in the elastically stressed parts of the cross section. These
contour lines at each point x,y give the direction of the resultant elastic
shearing stress t. It is clear that the elastic stress lines at their inter
sections with the boundary curves of the plastic regions must satisfy a
further condition: they must suffer no break on passing these points.
The shearing stress r must have the same value = k, regardless of whether
this value is approached from the plastic side or from the elastic side. A
break in the stress lines would be permissible only if the bounding curve
of the plastic region bisects the angle of the stress lines, in which case,
however, the vector of the resultant shearing stress r at the limiting line
must rotate through a finite angle. After the occurrence of an infinites
imal permanent deformation, no reason can be given for such finite
rotation. Therefore both branches of a stress line must be tangent to
each other at their intersection with the bounding curve of the plastic
1
Cf. Eq. (35-7).
PLASTIC TORSION 497

region. From this it follows that, if one chooses at a certain point of a


limiting curve F = F\, the ordinates F\ and F, of the elastic and the plastic
stress function, must satisfy at all points of the limiting curve the condi
tion F = Fi. It will be recognized that for a given angular twist 8 of
the bar the limiting curves of the plastic area of the cross section are the
projections of those curves on the plane of the cross section, along which
the elastic stress surface Fi just touches the plastic stress surface F.
From the known properties of the surfaces F and Fi the following experi
mental representation of stress distribution in a twisted bar, after the yield
point has been reached, may be determined: A horizontally placed piece of
cardboard having the geometrical shape of the cross section is covered
with sand so as to form under gravity a sloping surface or "heap" over
the cross section. From this heap a negative or "hollow" may be made
or a roof under a constant slope may be erected
according to this surface above the boundary of
the section. If the plane base of this roof is
closed with a stretched membrane and this mem
brane is loaded by pressure, parts of the mem-
i mi j. l ii_ r c it. c
brane will touch the surface of the roof erected
i jFlO. 86-9. Membrane and
8urface of constant Biope.
above the section (Fig. 35-9) when the pressure
reaches a certain value. The free parts of the membrane and those rest
ing on the sloping surface together form the stress surface for the plastic
ally deformed bar in torsion. Under those parts of the membrane touch
ing the sloping surface, the metal yields, while at those corresponding
to the free parts of the membrane the metal remains elastic.
Suppose that F(x,y) represents the ordinate of the stress surface at an
arbitrary point x,y, which may be chosen now either in the elastic or a
plastic region of the cross section of the partially yielded bar.
The following rules stated for the elastic torsion still hold true after
the yield point has been reached: the resulting shearing stress at an
arbitrary point x,y of the cross section is the gradient [grad F(x,y)] of the
stress surface F(x,y), the twisting moment M of the bar is equal to twice
the volume M = 2FJf(x,y) • dx dy, enclosed by this surface. The con
tour lines F(x,y) = const of the stress surface F(x,y) are the stress lines
of the cross section of the plastically twisted bar.1
1 If represents the angular twist at an instant at which the yield stress k (the
0o
yield stress in pure shear) is reached at one or more points at the edge of the cross
section, this angle is determined from the condition that at this point the maximum
value of the gradient of the corresponding elastic stress surface Fi is equal to

(grad Fi)m»i = k .

Then the pressure p0 at which the membrane first rests on the sloping surface has the
same ratio to an arbitrary pressure pi > p0 as the corresponding angular twist $0 has
498 THEORY OF FLOW AND FRACTURE OF SOLIDS

The construction of the contour lines of the plastic stress surface is very much facili
tated by some well-known properties of the surfaces of constant maximum slope.1 If
the edge of the cross section of a twisted bar is formed of straight lines or circular arcs,
the plastic stress surface forms a "roof" over the cross section consisting of planes or
portions of circular cones having the same slope.
The mode of penetration of the plastic areas into the inside of a twisted bar with a
rectangular cross section is shown in Figs. 35-10 and 35-11. The elastic area shrinks
gradually until finally it consists only of thin strips. The
middle lines of these strips are represented in the ground
plane of the stress surface by the ridge and oblique edges of
the "roof" erected on the rectangle. The stress surface
for plastically twisted strap iron is a narrow ridge, the peak
of the ridge being rounded off by a parabolic cylinder.
On the assumption that deformations are independent
of yield stresses, it follows, from this schematic represen
Fig. 35-10. Fig. 35-11. tation, that the values of the twisting moment approach
Fios. 35-10 and 35-11. asymptotically a certain value given by twice the space
Shaded areas indicate enclosed by the plastic stress surface.
plastic regions in cross
section of twisted bar.
The condition in which all parts of a twisted bar yield
may be designated as the completely plastic state.
For the completely plastic state the twisting moment of the bar may be easily calcu
lated. For example, taking a circular bar having a diameter 2a, the plastic stress
surface is a circular cone having the equation

F - Jb(r
- o) . (35-18)

In this k is the yield stress for pure shear. For F r = 0, = — ka, and the cone has a
height h = ka and a volume V = ra'h/S. According to our analogy the twisting
moment acting on the bar is equal to twice the volume of the cone, or

2*ka'
M = 2V (35-19)

For a bar having a cross section of the shape of an equilateral triangle, the plastic
stress surface is a three-sided pyramid. Each side has a slope equal to dF/&n = k,
since the shearing stress is equal to the slope of the stress surface. Therefore, the

to 0|. The unit angular twist of the bar which corresponds to given plastic areas
fli

indicated by the areas where the membrane touches the "roof") therefore
is

(as

= •.£-«.
«,

Po

Regarding the properties of surfaces of constant slope cf. F. Schilling, in Z.


1

angew. Math. Mechanik, vol. no. p. 197. It sufficient here to make few
3,

3,

is

remarks: The contour lines of a surface of constant maximum slope are equidistant
curves. The contour lines of the plastic stress surface may be obtained the normals
if

to the boundary curve are constructed and points having the same distance from the
edge are connected. In order to make a model of surface of constant slope a vertical
a

cylinder
is,

according to Schilling, erected on the evolute of the edge curve. strip


A

of paper cut obliquely so that its angle equal to the angle of slope wound around
is

is

this cylinder. The oblique straight line describes the sloping surface.
PLASTIC TORSION 499

pyramid has the height h = ka/2 \/3- The area of the base is y/Za*/4, its volume
V = a*h/4 -\/3, and the twisting moment when the bar is completely plastic is equal
to twice the volume or

For a bar with a rectangular cross section having the sides a and 6 (a is less than 6)
the moment in the completely plastic state is

M - *f + -
(35-21)

35-3. Apparatus for Experimental Determination of Stress Distribu


tion. The foregoing remarks permit the study by means of a special
apparatus of the manner in which plastic flow penetrates into a twisted
bar of steel. The shape of the stress function on which the stress distribu
tion depends can be experimentally determined with the help of a thin
rubber membrane first uniformly stretched in its plane and then loaded by
lateral pressure.1
The apparatus consists of an aluminum disk of 12 in. diameter, upon
which a uniformly stretched, thin rubber sheet was fastened by means
of an aluminum ring and screws. The rubber membrane was partly
supported along the circumference, but a central circular portion of it was
completely free. In the empty space closed by the membrane, pressure
could be produced by means of a tire pump (see Fig. 35-12). On the
aluminum disk a second aluminum disk could be fastened (d) which con
tained the flat "roof" described in the preceding section. This disk can
be seen on the right in Fig. 35-12. One of the roofs was made of glass (in
which case it had the advantage of being transparent and thus allowing
direct observation of the distorted rubber surface) and the other of brass
plates accurately machined. After the disk containing the flat roof was
fastened on the base, pressure was applied and the membrane allowed to
bulge out, until it partly or nearly completely covered the flat surfaces of
1 For the case of elastic torsion the soap-film analogy was first used by Anthes
(Doctor's Dissertation, Dresden, 1912), who photographed the picture of a rectangular
network of lines, which was reflected by the soap film produced over holes in a metal
sheet. Another method also using soap films was proposed by A. A. Griffith and
G. I. Taylor, Proc. 1st Intern. Congr. Applied Mechanics, p. 39, Delft, 1924, who
measured the deflections of such films by an electric contact method. An apparatus
utilizing a uniformly stretched rubber sheet, over which a pressure difference could
be produced by evacuating the air below the membrane, was demonstrated to the
author in 1928 by Prof. Enger at the University of Illinois. The principles used in
the latter apparatus were adapted in the one described above and constructed at the
suggestion of the author by H. Friedman at the Westinghouse Research Laboratories
in 1929, to whom the author is much indebted for carrying out all the tests referred to
above and for photographing the sand heaps.
500 THEORY OF FLOW AND FRACTURE OF SOLIDS

the metal roof. To make visible the boundaries of the plastic parts in a
cross section of a twisted bar (c/. under Sec. 35-2, page 49) the rubber
sheet was covered with a thin layer of white powder and the roof with a
thin film of oil, both put on before the pressure was applied. After the
latter was applied parts of the inflated membrane came in contact with
the flat surfaces of the roof and the white powder adhered to the roof,
where the rubber sheet rested on it. Thus white areas of the adhering

Fio. 35-12. Apparatus for experimental demonstration of stress distribution in plastic


torsion. An aluminum disk serving as base is covered by rubber membrane 6; c, aluminum
ring clamping membrane; d, disk containing constant slope surface; c, clamps;/, connection
to pressure-measuring device.

powder disclosed the shape of the common surfaces of contact or the


plastic parts of the cross section. These parts appear white in the photo
graphs taken of the hollow side of the roof after it had been removed.
A series of tests showing the progress of the plastic regions in a twisted
bar with a square cross section can be seen in the photographs, Figs.
35-13 to 35-15. The figures below indicate the pressure; the angle of
twist would increase in the same proportion as the pressure. From Figs.
35-13 to 35-15 it may be seen that the white areas are first formed in the
middle of the sides of the square, where the shearing stress first reaches
the yield point according to the theory of elastic torsion. With increas
ing twist or moment these white areas, representing the portions which
become gradually plastic in the cross section, grow inward. Finally,
under sufficiently large pressure only a narrow dark cross appears (Fig. 35-
15), showing that the elastic portion in the cross section has been reduced
PLASTIC TORSION 501

Fio. 35-15.
Fios. 35-13 to 35-15. Development of plastic regions in a twisted bar of square cross
section. Elastic areas appear dark, plastic areas bright. (Pressures p acting on rubber
membrane were proportional to 1.5/1.75/2.5.)
502 THEORY OF FLOW AND FRACTURE OF SOLIDS

practically to two narrow strips, crossing at right angles and following the
course of the two diagonals of the square or the projections of the corre
sponding four edges of the roof.
The shape of the plastic stress function F(x,y) or of the stress surface

Flo. 35-16. Sand heaps produced over rectangles showing constant slope surfaces.

for the yielding of the whole bar can be demonstrated by


case of complete
sand heaps covering the figure of the cross section of the twisted bar.
Such sand heaps were produced and can be seen in some photographs
(Figs. 35-16 to 35-19) taken for the square, rectangular, an equilateral
triangular, an elliptical, and two circular cross sections containing grooves
PLASTIC TORSION 503

Fio. 35-17. Sand heap over equilateral triangle.

Fig. 35-18. Sand heap over an ellipse.

a b
Fig. 35-19. Sand heaps over areas bounded by two circular arcs.

of semicircular form. This last case would correspond somewhat to a


shaft with a key way of semicircular cross section.1
1 The
case of a twisted shaft with a keyway was also the subject of similar testa as
those mentioned above in a paper of E. G. Coker, Elasticity and Plasticity, Proc.
Inst. Mech. Engrs., London, p. 897, November, 1926 (Thomas Hawksley lecture), to
which special reference is made. In this paper Coker has applied the membrane and
sand-heap analysis for the experimental determination of stress distributions in twisted
shafts containing a keyway.
504 THEORY OF FLOW AND FRACTURE OF SOLIDS

The models of the plastic stress surfaces for complete yielding can be produced
experimentally by placing plates of metal cut according to the shapes of the cross
sections of a twisted bar on an iron frame and by covering it and the plates with a
layer of loose sand. The iron frame is supported in a horizontal position by four
strings and may be lifted by a crank mechanism (Figs. 35-20 and 35-21). By gently

Fio. 35-20. Sand-heap apparatus

Fig. 35-21. Sand-heap models.

raising the frame carrying the plates the sand models may be produced in almost
perfect form. A crushed zirconium oxide powder (beach sand found in Australia)
served well for this purpose.1
1 "The Phenomenon
of Slip in Plastic Materials," Edgar Marburg lecture, ASTM
Chicago meeting, 1931. An experimental method for determining the distribution of
stress in elastic torsion has been proposed by A. Piccard and W. Tohner utilizing a
meniscus of a fluid. (C/. "Stodola Festschrift," Verlag Julius Springer, Berlin, 1929.)
E. Trefftz was the first to solve a problem of the gradual spreading of a plastic
zone from a fillet in the section of an elastically twisted bar causing a concentration of
the shearing stresses in the vicinity of the fillet. He considered an angle iron and
determined the shape of the yielded region at the inner reentrant corner of the iron
PLASTIC TORSION 505

It is noted that the traces or curves in which the boundary surface between the
elastic and the plastic zones intersects the planes of the cross sections of twisted bars
with the exception of the circular section traverse the contour lines F = const of the
plastic stress function or the stress lines within the section. The latter curves — it
may be mentioned — are one set of the "characteristics" of the partial differential
equation of hyperbolic type [Eq. (35-15)] of the plastic stress function F.1 We may
conclude that an elastic-plastic inner boundary surface in general intersects the
"characteristic" curves or surfaces in partly yielded bodies and may not be assumed
to coincide with the traces of these surfaces, as has been postulated erroneously by
investigators discussing such cases, for example, under the states of plane strain.
35-4. Yielding in Bars Having Holes. In the preceding text no reference has yet
been made to the cases of torsion in which the cross section of the twisted bar contains
holes. Mathematically such cases are referred to as double or multiple connected
areas. In these cases certain additional conditions must be satisfied when the solu
tions for the two stress functions are established. M. A. Sadowsky* pointed out,
when discussing such cases for the complete yielding of twisted bars having holes, that
the sand-heap analogy may still be applied with a slight modification.
The sand heap representing the plastic stress surface of a hollow cylindrical bar of
equal wall thickness twisted around its axis is a circular plateau having conical slopes.
In fact, the torque of such a hollow cylindrical bar is determined by that part of the
volume or weight of a heap of sand piled up in the shape of a circular cone above the
outer circle of radius a which reaches up to the inner circle of the hollow section of
radius at. The torque M under which the hollow section yields completely is given
by the proportion

= (35-22)
w. w.

using the method of conformal mapping developed in the theory of functions of a com
plex variable (Z. angew. Math. Merhanik, vol. 5, p. 64, 1925).
Plastic stress surfaces for cylindrical bars of curvilinear cross sections (elliptic
cylinder) were discussed by L. S. Leibenson ("Elements of the Mathematical Theory
of Plasticity," 1943, Moscow, translated by Brown University, Providence, p. 135,
1917).
The transition from the elastic to the plastic state in a cylinder whose cross section
approaches an ellipse and in which an elastic nucleus of elliptic shape forms was
investigated by W. W. Sokoi.ovsky (J. Applied Math. Mechanics, Acad. Sciences,
U.S.S.R., vol. 6, p. 241, 1942, translated by Brown University, Providence, 1946);
the case of partial yielding in a bar of square cross section by L. A. Galin (ibid.,
vol. 8, No. 4, p. 307, 1944).
The warping of the cross sections in prismatic bars twisted beyond the yield point
was discussed by P. Hodge (Office of Naval Research Tech. Rept. No. 18, May, 1948,
Brown University, Providence).
The plastic torsion problem may be also treated more generally by assuming
a strain hardening function Tom = /(Tool) of the material, e.g., a power function
Toot = cToct" (author's paper in "Theodore von Karman Anniversary Volume,"
1941).
1 For the theory of characteristics, see Chap. 37 on plane stress.
1 An Extension of the Sand Heap Analogy in Plastic Torsion Applicable to Cross
Sections Having One or More Holes, J. Applied Mechanics, June meeting, 1941.
506 THEORY OF FLOW AND FRACTURE OF SOLIDS

in which M, is the torque of a solid cylinder of radius o and may be determined by


the ratio of the weights of the lower portion of the cone W and of the entire sand
cone W, which can be piled up above the base circle of radius a.
The plastic stress surface of a round bar having an eccentric cylindrical cavity is

(a) (b) (c)


Flo. 35-22. Plastic torsion. Stress surfaces for bars having cross sections with holes.

Fio. 35-23. Fig. 35-24.


Flos. 35-23 and 35-24. Metal frames for producing sand heaps.

Fio. 35-25. Fio. 35-26.


Flos. 35-25 and 35-26. Sand piles representing plastic stress function for complete yielding
of a round and a square bar having holes.

represented by its contour lines in Fig. 35-22<z and may be produced in a sand heap by
the model apparatus shown on the left of Figs. 35-23 and 35-24, consisting of a circular
metal disk in which a hollow metal cylinder can slide through a fitting hole. Accord
ing to Sadowsky, the sand heap corresponding to a twisted round bar having an eccen
trically located circular hole is obtained after one has raised the sliding metal tulie to
the proper height before covering the area outside the "hole" with sand. If the
metal tube is not raised sufficiently high, less sand will pile up as required across the
PLASTIC TORSION 507

narrowest portion of the ring-shaped cross section because a ridge will form in the sand
pile (having a positive and a negative slope — a fact which contradicts the mechanical
requirement that the shearing stresses in this region shall have the same sign, since
the slopes of the stress surface F represent the shearing stresses). If the metal tube,
on the other hand, was raised too high, the sand pile will not satisfy the "boundary
condition" along the inner edge of the cross section, where the latter should form a
horizontal contour line of the stress surface F. The correct shape of it is illustrated

Fig. 36-29.

Figs. 35-27 to 35-29. Distributions of stress in the corner region of a hollow rectangular
cross section of a bar subjected to torsion. (After F. S. Shaw.) Represented by curves of
constant maximum shearing stress (grad F = const).
Fig. 35-27. Elastic stage just when yielding starts in fillet (0 = 0o).
Fio. 35-28. First plastic stage (9 = 1.759o).
Fig. 35-29. Second plastic stage (9 = 1.8759o).

by a sand model reproduced in Fig. 35-25 corresponding to the contour-line picture of


Fig. 35-22a and to a mass of sand whose surfaces consist of two intersecting cones of
opposite slopes. The "torque" may be obtained by weighing this sand pile, after
having added to the quantity of sand shown in Fig. 35-25 as much sand as will fill up
the empty space in the metal cylinder above the ground plate to the rim of the sliding
metal tube. Two more examples may be seen in Figs. 35-226 and c, 35-24, and 35-26,
showing a metal frame for producing the effect of a square hole and the shapes of the
stress surfaces.
The Australian engineer F. S. Shaw1 has admirably extended the theory of partial
1 "The Torsion of Solid and Hollow Prisms in the Elastic and Plastic Range by

Relaxation Methods," Australian Council for Aeronautics, Report ACA-11, Novem


ber, 1944, Melbourne, Australia, 38 pp.
508 THEORY OF FLOW AND FRACTURE OF SOLIDS

yielding to several examples of the spreading of plastic zones around the fillets in an
I beam of standard profile carrying a torsion moment, in a hollow splined shaft having
projecting ribs, in a bar having a hollow rectangular cross section with fillets at the
reentrant inner corners, and in a round bar having an eccentric cylindrical cavity, all
subjected to torsion. Shaw constructed the elastic and plastic stress functions by
the method of finite differences after replacing the differential equations of these two
functions by equations containing the finite differences of their values at a large num
ber of points and after solving these linear equations by the "relaxation" method.
The distributions of the shearing stresses which he obtained in the corner region of a
hollow rectangular cross section are reproduced by means of the contour lines of a
surface whose ordinates were equal to rm, i.e., equal to the slopes of the stress func
tions or equal to grad F. In the elastic region along these curves rm < k and in the
plastic regions rm = k. In the region near a fillet at a reentrant corner of the sec
tion according to the theory of elastic torsion a concentration of the stresses occurs,
which depends on the proportions of the dimensions of the section. The bar may start
to yield either at the center point of one of the outer straight edges or at the fillet of
the inner edges. The proportions in Fig. 35-27 were such that yielding started at the
fillet. The shaded portions in Figs. 35-28 and 35-29 are the plastic regions. The wall
thickness of the hollow section being a small fraction of the outer dimensions L and
2L, the stresses away from the corner are soon distributed according to a cylindrical
surface whose contour lines become parallel to the straight edges of the section. If 0»
represents the angle of twist of the bar when the yield stress is just reached in the
reentrant corners of the section, the angles of twist corresponding to the "first" and
"second" stages of stressing which are illustrated in Figs. 35-28 and 35-29 are chosen
equal to 1.75»o and 1.8750O.1

36-5. Partial Yielding in a Twisted Shaft of Variable Cross Section.


In the theory of elasticity two stress functions have been introduced for
the treatment of the torsion problem of bodies of revolution or of shafts
under the action of a torque and having diameters which vary with the
1 The inner edge of the membrane surface representing the clastic stress function
must be raised gradually above the base plane of the cross section as the torque in the
bar (the lateral uniform "pressure" under which the membrane is deflected) increases.
The height to which it must be raised is determined from the condition that the
integral of the rotation in the plane vector field of the resultant shearing stress vectors
t taken over the area of the cross section must become equal to the line integral of t
taken around the outer and the inner edge of the cross section (theorem by Stokes).
When this height has reached a certain value, grad F will just become equal to k at
the fillet. Beyond the corresponding position of the inner edge of the membrane
two additional surfaces must be considered of constant slopes, one of which is to be
erected above the outer and the other below the raised inner edge of the cross section,
the latter in a direction sloping downward. The latter surface (inner roof) of con
stant slope (grad F = const) is assumed floating and its correct position has to be
determined by making use of Stokes' theorem. The membrane will first come in
contact with the inner roof defining the first plastic region in the vicinity of the reen
trant corners, and subsequently it will also become tangent to and touch the outer
roof when the torque has further increased and additional plastic regions will form
along the outer edge of the cross section. Two such stages of twisting are seen in
Figs. 35-28 and 35-29.
PLASTIC TORSION 509

cylindrical coordinate z measured in the direction of the axis of revolu


tion.1 Let r and z be the cylindrical coordinates of a point within the
body in radial and axial direction. It is easy to see that in a shaft of
variable diameters subjected to a twisting moment only two components
of stress act, namely, the shearing stresses t, and r, whose arrows2 are
shown in the faces of a small element in Fig. 35-30, furthermore that the
components of displacement in the radial and axial direction vanish at
every point, leaving only a nonvanishing tan
gential displacement v by means of which the
unit shears are expressed as follows: jj •
_
y'~Tr
dv _ v
= ?1
(35-23)
r dz'
Hence

*-•£-;)•
dv
(35-24)

The equilibrium condition of these two compo


nents of stress:

l (A,) +
|
is satisfied, if one assumes that
<A0 = 0 (35-25)
Fig. 36-30.

rT' =~Tz and r-T, = (35-26)


dr

The compatibility of Eqs. (35-24) requires that

3> _33tf. S*± = n


dr2 r dr dz' (35-27)

Since on the boundary curve r = f(z) in a plane section through the shaft
passing through its axis the resultant shearing stress r = \Zrr2 + t22
must be tangential to this curve, we have

dr
dz
Tr
T,
u dr - tt dz = if I* dr +
|J J rfz = 0 (35-28)

from which it follows that the boundary value of the stress function

1 See
FttPPL, A., "Vorlesungen iiber Technische Moehanik," Vol. 5, p. 183, B. G.
Teubner, Leipzig, 1922. Timoshenko, S., "Theory of Elasticity," p. 276, McGraw-
Hill Book Company, Inc., New York, 1934.
* Instead of writing rr, and t,< we use the abbreviated symbols t, and t,.
510 THEORY OF FLOW AND FRACTURE OF SOLIDS

RELATIVE VALUES
OF SHEARING STRESS T
48
45

AXIS OF SHAFT
Fig. 35-31. Partial yielding near fillet represented by means of the curves of constant
resultant shearing stress (r = Vv + t«* = const) in section of a cylindrical shaft with a
collar twisted around its axis. (After Eddy and Shaw.)
<t>

=
<f>

= = const.1 Assuming along the axis = the torque

is
0)
(r
0

4>a

found equal to

M - 2r °
rr* dr = 2rd>a (35-29)
J

R. P. Eddy and F. S. Shaw2 have extended this theory of the elastic


torsion of body of revolution to the case of partial yielding. Similarly,
a

second stress function may be introduced by letting = nf/, Gr* — = —


1
A

v
$

dr dz
M
Gr' — = —
<t>

d<t>
<p

After eliminating here one obtains for


"*" "*" ' (35-28o)


dr' dr 3«'
r

The surfaces = v/r = const represent surfaces within the body having the same
^

constant angle of twist.


Numerical Solution of Elasto-plastic Torsion of a Shaft of Rotational Symmetry,
1

J. Applied Mechanics, ASME meeting, 1948.


f
PLASTIC TORSION 511

as in the case of a twisted prism treated in Sec. 35-2, suppose that in cer
tain portions of the body the yield stress is reached. Looking upon the
plane section passing through the axis the components of shearing stress
tt and t, have a resultant r = vV V
+ and since no normal stresses in
the directions perpendicular to the meridian planes act, the condition of
plasticity requires that in the yielded portions of the body the resultant
stress
t = vV + t,2 = k = const . (35-30)

Since the stress components rr and t, must here also satisfy Eqs. (35-25)

<tt>
and (35-26), there results for the plastic stress function from Eqs.
(35-30) in the yielded portions of the body the differential equation

This again "floating" surface of given shape whose ordinates on the


is

<tt>

boundary curve must be made equal to = <£„ = const. In the elastic


region < Ax2. The main part of this boundary value problem for the
t

<t>

elastic stress function satisfying Eq. (35-27) consists in the construction


<$>

of these surfaces for number of sufficiently large torques which will


a

become tangent to the plastic stress surface satisfying Eq. (35-31). The
curves along which the elastic stress surfaces touch the plastic surface
then determine the inner boundaries of the plastic regions in the cross
section. Eddy and Shaw worked out the case of cylindrical shaft with
a

collar by numerical method similar to the one used in the paper by


a
a

Shaw quoted on page 507 (Fig. 35-31).


CHAPTER 36

THE FLOW LAYERS IN TWISTED BARS OF MILD STEEL.


EFFECT OF GROOVES AND HOLES

36-1. Flow Layers. Torsion Tests with Steel Bars. The position of
the flow or slip layers in the plastically deformed parts of a twisted bar
may be predicted with the help of the sand-heap analogy. For a soft
metal, such as mild steel, the slip layers coincide approximately with the
surfaces of maximum shear or of maximum shearing dis
placements. These slip layers are therefore approximately
perpendicular to each other. At an arbitrary point inside
of a twisted bar, one surface of maximum shearing stress
coincides continuously with the plane of the cross section.
The other surface of maximum shear is parallel to the axis
, I i of the bar, i.e., perpendicular to the cross section. The
traces of the second system of slip layers must remain per
^t"^
Flq.
pendicular to the stress lines of the plastic stress function.
36-1.
Orientation After the yield stress has been passed in a twisted iron
of flow layers
bar, very regular layers or markings may actually be shown
in a steel bar
twisted above to exist.1 In these, the iron is apparently deformed much
the plastic more than in the neighboring layers. These layers may sub
limit.
sequently be made visible in the cross section of the bar
by means of Fry's etching method. In such etched sections the flow
layers appear as dark strips and lines. A schematic sketch of their

Angle of Twistper Inch


Flo. 36-2. Torque-twist diagram of a mild Flo. 36-3. Lines of greatest slope on roof.
steel showing flow or strain figures.

appearance in an oblique cross section of a bar with square cross sections


is shown in Fig. 36-1. A series of such etchings on soft-iron bars, using
1 The observations referred to above were published in a paper by W. Baoer and
the author, Die Vorgange nach der Uberschreitung der Fliessgrenze in verdrehten
Kisenstaben, Z. Ver. deut. Ing., No. 10, p. 317, Berlin, 1927; and in the Doctor's Dis
sertation by W. Bader, University of Gottingen, 1927.
512
FLOW LAYERS IN TWISTED BARS 513

Fry's method, are shown in Figs. 36-4 to 36-8 for circular cross sections, in
Figs. 36-9 to 36-16 for square and rectangular cross sections, and in Figs.
36-17 to 36-19 for triangular cross sections. In the etched cross sections
the traces of one of the two systems of slip layers are indicated. Traces of

Figs. 36-4 to 36-8. Development of flow layers in twisted steel bars. Diameter of bar,
17 mm. Angle of twist per unit of length, 6: Fig. 36-4, 0 = 0.007 deg; Fig. 36-5, 6 = 0.04
deg; Fig. 36-6, 6 = 0.22 deg; Fig. 36-7, 6 = 1.84 deg; Fig. 36-8, 6 = 3.02 deg.

Fio. 36-9. Fio. 36-10. Fio. 36-11.


8 = 0.38 deg. 6 = 0.68 deg. 6 = 0.9, deg.
Figs. 36-9 to 36-11. Fluidal structure in twisted bars with square cross section. Square
2 by 2 cm. 6, angle of twist per centimeter.

the second system of slip layers could, however, seldom be noted upon the
surface parallel to the generatrices of the cylindrical or prismatical bars
or in sections inclined to the axis of the bar.
In uniformly when the twist
a twisted steel bar the flow layers appear
ing moment reaches values corresponding to those along the horizontal
portion AB of the moment curve M = f(9), Fig. 36-2. The first flow
514 THEORY OF FLOW AND FRACTURE OF SOLIDS

Figb. 36-15 and 36-16.


Fios. 36-12 to 36-16. Flow layers in twisted steel bars of rectangular cross section.

Fio. 36-17. Fio. 36-18. Fio. 36-19.


Flos. 36-17 to 36-19. Flow layers in twisted steel bars of triangular cross section.
FLOW LAYERS IN TWISTED BARS 515

layers occur at those values of twisting moment at which the outermost


part of the cross section has already been plastically deformed and the
moment curve begins to bend over into the horizontal branch. With
increasing twist, new layers form beside the old ones, while the latter
become longer and thicker. The flow layers appear in the etched cross
sections of the photographs as wedge-shaped areas, the points of the
wedges pointing toward the less stressed parts of the cross section (Fig.
36-9).
As a rule, it may be concluded that the black flow lines in the cross
sections run mainly perpendicular to the stress lines (the projections of
the contour lines of the plastic stress surface). Moreover, these flow
lines always remain perpendicular to the edge of the cross section. Thus
they run in the direction of the projections of the lines along which water
will run off the roof representing the plastic stress surface (Fig. 36-3).
With the more severe deformations the black strips finally cover the
whole cross section. In the etching of the square cross section of
Fig. 36-11, which corresponds to the most severely twisted bar,
the elastic areas may be recognized only as bright areas along the
diagonals of the square. In other words, the four ridges of the
stress surface F, which is here a four-sided pyramid, are projected
on these diagonals.

36-2. Structure Due to Cooling and Fluidal Structure. The similarity of the
crystalline structure in cast-metal ingots1 with the photographs here given of the flow
layers in twisted bars, especially for the circular
and square cross section, suggests the necessity B f
for an investigation of whether or not a corre
spondence exists between the formation of the
flow layers in twisted steel bars and the figures
which one may observe in the macrostructure of
rolled material as the traces of the ingot structure
o A
in the cross sections.
For this purpose a square bar, having the cross Fio. 36-20. Fio. 36-21.
Fias. 36-20 and 36-21. Structure
section EFGH (Fig. 36-20), was cut out from a
of ingot and flow structure in steel
rolled-iron bar having an initial cross section bars.
A BCD. Moreover, a bar with a cross section
made up of four circular arcs (Fig. 36-21) was machined from a bar A BCD as shown
in Fig. 36-21. The flow structure of these bars, when twisted, was compared with
that obtained previously. In Figs. 36-20 and 36-21, the structure produced by
crystallization in the ingot, which is not changed by rolling, is represented by the
thin lines on the left half of the figure, while on the right half of both figures the
normal position of the slip or flow layers produced by torsion is given. The flow
1Remarkable examples are contained in the books by J. Czochralski, "Moderne
Metallkunde in Theorie und Praxis," pp. 98, 103, 139, Berlin, 1924; and by P. Ober-
hoffer, "Das technische Eisen," 2d ed., p. 291, Berlin, 1925; cf also E. Seidl and
E. Schiebold, Das Verhalten inhomogener Aluminium-Querblockchen beim Kalt-
walzen, Z. Metallkunde, vol. 17, pp. 225/f., 1925.
516 THEORY OF FLOW AND FRACTURE OF SOLIDS

layers, as found by the etched cross section of the twisted bars (Figs. 36-22 to 36-25),
correspond to the actual position of the applied stresses. A marked influence of the
structure produced by rolling upon the observed shape of the flow layers therefore
does not appear to exist.
For comparison the behavior of a brittle cast metal, in this case a cast-zinc bar
(Fig. 36-26), which has a very definite structure produced by cooling, was investigated.
A square bar of zinc was cast in an iron mold and then twisted in a torsion machine.
The bar broke in torsion in the manner peculiar to brittle materials and under rela
tively small stresses by exceeding the tensile strength along a surface inclined at

Fio. 36-22. Fig. 3,-23.


Fios. 36-22 and 36-23. Flow layers in twisted steel bars cut from a bar according to
Fig. 36-20.

Fig. 36-24. Fig. 36-25.


Fios. 36-24 and 36-25. Flow layers in twisted steel bars cut from a bar according to
Fig. 36-21.

45 deg to the axis of the bar. In the fractured surfaces and in the etching of the cross
section the ray-shaped texture of the microstructure produced by cooling of the cast
ing was evident, as shown by Fig. 36-26.
Further evidence that the fibrous structure produced by rolling has no marked
influence on the position of the flow layers is afforded by the torsion tests of bars with
triangular and irregular cross sections. All these bars were machined from bars with
square or rectangular cross sections; in the etched cross sections the dark square-
shaped zones may often be plainly recognized, while the flow layers run in the direc
tions predetermined by the torsion stress field and correspond to the rules laid out
above.
A systematic difference in the position of the dark strips in the etched cross sections
from that of the lines of maximum slope of the stress surface must be noted, however.
The pointed ends of the flow lines appear to bend over and to follow the narrow- bright
FLOW LAYERS IN TWISTED BARS 517

strips in the etchings. It appears as if the lines of maximum slope of the plastic stress
surface, after severe torsional distortion of the test piece, tend to approach again the
curves of maximum slope of an elastic stress surface. Probably this phenomenon
depends on the fact that the strengthening of the steel or cold working begins as a
result of preceding plastic deformation, which fact was neglected in the simple theory.1

Fio. 36-26. Structure of cast zinc showing crystallization due to rapid cooling.

36-3. Longitudinal Groove with Semicircular Cross Section. For


the case of a bar of circular cross section in torsion, having a longitudinal
groove of semicircular cross section, qualitative estimates of the positions
of the flow layers may easily be made. We shall first assume that the
radius a of the groove is small in comparison with the diameter d of the
bar. It is then permissible to replace the boundary circle of the cross
section by its tangent. The disturbances in the direction of the stress
lines may then be determined under the assumption that, at a large
distance from the groove, a constant shearing stress acts.
The question is now to determine the stress distribution in a body
subjected to pure shear in a direction parallel to one edge and having a
small groove with a semicircular cross section (Fig. 36-27). For the case
of pure elastic shear, the stress lines in the neighborhood of the groove
may easily be determined with the help of the membrane analogy. One
has only to think of a thin membrane attached along an edge ABCDE
1 Various papers by M. J. Seiole consider the plastic deformation of a twisted bar.
Cf. especially Quelques particularitfe thobriques et experimentales de la torsion des
barreaux a section non-eirculaire, Rev. ind. mintrale, p. 557, 1925. In the above-
mentioned work Seigle used the phenomenon of recrystallization to observe the plastic
regions. Relative to further studies see also papers by Seigle and Cretin on torsion
and tension in Gtnie Civil.
518 THEORY OF FLOW AND FRACTURE OF SOLIDS

(Fig. 36-27) of the body, this membrane being stretched in a plane


inclined to the plane of the cross section. The slope of the inclined plane
is a measure of the amount of shearing stress in the
undisturbed part of the stress field at a large distance from
the groove. The contour lines of the stretched membrane
near the edge are the stress lines of the region subjected
to pure shear (Fig. 36-28). These lines are closest to
gether in the neighborhood of the point C at the edge of
the groove, and at a large distance from it they become
Fig. 36-27. more and more nearly parallel to the edge. The shearing
Semicircular stress t which exists at a given point P is equal to the slope
groove on of the elastic stress surface F.1
straight edge.
As the inclination of the plane of the thin membrane is
increased, all shearing stresses are increased until the stress at the point C
reaches the yield point. Then the membrane must be inclined in a plane

Fio. 36-28. Membrane Fio. 36-29. Mem Fio. 36-30.


Stress lines around
near semicircular notch brane and cone.
and straight edge. semicircular groove
in straight edge.
Shaded area repre
sents plastic region.

F =kx/2 with respect to the plane of the cross section. In order to


determine the limits of the plastic area we must think of a sloping sur-
1 By means of the potential theory the stress function F may easily be determined.
Using polar coordinates r and a

F
"(?-') cos a .

By forming the derivatives of this surface the maximum shearing stress t^m, is found
to occur at the edge of the groove at the point C, where r = a and a = 0 and is
Tmu = 2c, i.e., twice as large as the shearing stress t = c = const at a large distance
from the groove. Since when i
approaches infinity F = — cx and the shearing stress
r = —dF/dx = c, the picture of the contour lines in this case is exactly the same as
that of the streamlines of a fluid which circulates around a circular cylinder, the liquid
moving perpendicularly to the axis of the cylinder. (Only one-half of the picture of
the streamlines around the cylinder has to be considered.)
FLOW LAYERS IN TWISTED BARS 519

face in form of a cone, erected, as shown in Fig. 36-29, above the circle of
the radius r = a. If the slope of the membrane at C becomes larger than
k/2, certain neighboring parts of the membrane will lie upon the cone.
These parts, which have been shaded in Fig. 36-30, represent the plastic
area of the cross section. Of especial interest here is the limiting case
when the shearing stress r = c approaches the value of the yield stress

t = k f or pure shear. Calculation1 showed that when t approaches k and

^^r
Flo. 36-31. Apparatus for experimental demonstration of spreading of plastic region from
semicircular notch or cylindrical hole. The dark block at the left is a micarta cone, which
can be fastened to the frame by means of the four screws shown.

for x = «) the plastic region becomes bounded by two parallel lines. The
plastic region becomes a parallel strip having a finite width b which was
found equal to

This leads to the interesting conclusion that, in a body stressed by pure


shear and having a small semicylindrical notch, with the increasing stress
only a thin layer of plastic material will be formed. When the shearing
stress approaches the yield stress of the material, this plastic layer
extends indefinitely and has a width 2/V times the diameter d = 2a of
the notch. It is not difficult to obtain the solution of the problem dis
cussed above by means of a mechanical apparatus. Figure 36-31 shows
an apparatus constructed for this purpose, consisting essentially of a

Proc. 2d Intern. Congr, Applied Mechanics, p. 337, Zurich, 1926.


1 See The stress
distribution around a hole after the yield point has been reached, has been treated by
E. Trefftz in Z. angew. Math. Mechanik, vol. 5, p. 64, 1925.
520 THEORY OF FLOW AND FRACTURE OF SOLIDS
FLOW LAYERS IN TWISTED BARS 521

sheet of rubber stretched in its plane on a rectangular frame of metal.


The plane of this frame could be tilted about one of its sides. The
rubber sheet could be brought by this movement into a position, in which
it partly touched and covered a cone attached to the frame of the appa
ratus in the manner indicated in Fig. 36-29. The contour lines of the
plastic region were made visible by the same method as used in the cases
of other cross sections already mentioned. The white areas
of an adhering powder on the dark surface of the cone (the
latter was machined from a piece of dark micarta) illus
trate nicely the shape of the plastic region. The spreading
of the plastic region from the edge of a semicircular notch
is shown in a series of photographs in Figs. 3G-32 to 36-35, Fig. 36-36.
taken by H. Friedman for various increasing angles of incli Plastic area
produced by
nation of the plane of the rubber sheet. The figures indi semicircular
cate that when the shearing stress in the undisturbed parts notch in re-
g i o n sub
of the stressed body approaches the yield stress for pure jected to pure
shear, the plastic region tends to take a more and more shear.

longitudinal shape.1
These considerations lead to a somewhat paradoxical conclusion; for
according to what has been stated previously, the presence of a small
longitudinal groove on the surface of a material strained by pure shear is
sufficient to cause the small plastic area at the
edge of the groove to extend deeply into the
material, if the shearing stress at this edge is
increased to the yield point (Fig. 36-36). More
over, from this it follows that it is not possible,
by increasing the shearing stress, to produce plas
Fig. 36-37. Plastic stress tic deformation elsewhere than in the narrow
function represented by flow area. The apparent discrepancy between
contour lines for corru
gated surface of bar. this conclusion and actual yield tests on bars is
easily explained by recalling one of the idealizing
assumptions on which it was based, namely, that the yield stress should
remain independent of deformation. On the other hand, the above con
clusions have been confirmed by numerous observations on the develop

1 Moreover, in the case of a corrugated surface the limits of the plastic areas may
be studied easily. If the edge of the cross section is approximately a wavy line, the
plastic stress function forms a roof, having grooves inclined to the edge and being
separated by wedge-shaped projecting edges, as shown by the contour lines of Fig. 36-
37. Since a stretched membrane can never touch these edges, it is seen at once that
such unevennesses of the surface of a twisted bar will give rise to oblong flow areas,
which under increasing stress will spread inward from the small grooves in a perpen
dicular direction to the mean edge of the cross section.
522 THEORY OF FLOW AND FRACTURE OF SOLIDS

ment of the flow layers and flow figures in plastic materials, which begin
to flow under a constant or a decreasing stress.
36-4. Cylindrical Hole. In the neighborhood of a small cylindrical
hole of diameter d in a region stressed by pure shear, two narrow flow
areas of width 2d/ir develop in an analogous way, if the shearing stress is
gradually increased to the yield point (Fig. 36-38). The stress distribu
tion in the neighborhood of a cylindrical hole in an infinite plate corre
sponds completely with the above-mentioned case of a semicylindrical
groove; the two cases are mathematically the same. At small holes,
notches, or grooves, the axes of which lie in the plane of principal shear,
under sufficiently high stresses, narrow flow areas
develop. The difference in the action of a small
hole, according to whether the strains are purely
Uyer^oduced'ty'cyir- elastic in the vicinity or whether plastic flow
drical hole in region sub- occurs (in the case of a material having a definite
jec e o pure s ear.
yield point) has already been shown by the tensile
or compression tests of specimens having holes which were discussed
in Sees. 18-3 and 18-4.
If, in the neighborhood of a hole, the strains are purely elastic, the dis
turbances in stress produced by the presence of the hole rapidly decrease
with distance from the hole. In a plastic material which begins to flow
under approximately constant stress, there results, however, in the case
of axial tension or compression, severe plastic deformation along two
planes, the deformation being quite marked even at a large distance from
the hole.
In soft-annealed copper test specimens it is usually not possible to
observe flow layers after plastic deformation has occurred. If, however,
the copper is subjected to severe cold working before the test, it acquires
a definite plastic limit and it is then possible to produce flow layers.
An essential condition for the production of the narrow flow layers in
plastically deformed materials seems to be that the materials possess a
definite yield point, i.e., a break in the ordinary stress-strain curve such
that the material yields under a constant stress or the load drops sud
denly, together with local concentration of plastic deformation because
of an increase in stress. A possible origin of such local concentrations of
stress is a small flaw or a slightly softer inclusion which has a somewhat
lower yield point than the neighboring parts.

Although no attempts seem to have been made as yet to analyze the plastic zones
forming around a small ellipsoidal cavity in an elastic body subjected to a state of
homogeneous triaxial stress at larger distances from the cavity, attention is called in
this connection to a remarkable paper by M. A. Sadowsky and E. Sternberg1 in which
1 StressConcentration around a Triaxial Ellipsoidal Cavity, J. Applied Mechanics,
ASME meeting, December, 1948.
FLOW LAYERS IN TWISTED BARS 523

these authors have developed the exact solution for the distribution of the stresses
around a cavity of triaxial ellipsoidal shape in elastic material for the case that the
body at infinity is in a state of uniform stress whose principal directions are parallel
to the axes of the ellipsoidal hole. Their solution was expressed in closed terms utiliz
ing the Jacobian elliptic functions, and formulas were derived in it for determining

Fio. 36-42. Fig. 36-43. Fig. 36-44.

Fig. 36-45. Fig. 36-46. Fig. 36-47.


Figs. 36-39 to 36-47. Fluidal structure in twisted steel bars having notches or grooves.
On right the contour lines of the plastic stress surface.

the concentration of the stresses caused by the ellipsoidal cavity. This general solu
tion includes the special case of such a cavity in a field of pure shear oi = a, oi = —a,
as = 0 assuming that two of the three principal axes of the ellipsoidal cavity are
parallel to the principal stresses o\ and at and as further special cases a cavity in the
shape of an elliptic cylinder and the spherical hole.
36-6. Bars with Longitudinal Grooves. In Figs. 36-39 to 36-56 are shown etchings
of twisted cylindrical bars having longitudinal grooves of different shapes and dimen
sions. In the flow figures found on the cross sections of a round bar with a small
524 THEORY OF FLOW AND FRACTURE OF SOLIDS

semicircular groove (Fig. 36-39), it will be noticed that three wide flow layers extend
from the groove. Figure 36-40 represents a similar round bar whose surface was,
however, given a mirrorlike polish before the test. Although the twisting moment
applied to this bar was about 10 per cent greater than that applied to the bar shown
in Fig. 36-39, only one flow figure was obtained near the groove.

Fio. 36-4S. Flo. 36-49. Flu. 36-50.

2>n*
1"-
- c}— ■■
Fig. 36-51. Fici. 36-52. Fra. 36-53.

Fig. 36-54. Fig. 36-55. Fig. 36-56.


Fias. 36-48 to 36-56. Fluidal structure in twisted steel bars having notches or grooves.
On right the contour lines of the plastic stress surface.

Figures 36-39, 36-42, 36-45, 36-48, 36-51, and 36-54 represent slightly twisted bars;
Figs. 36-40, 36-43, 36-46, 36-49, 36-52, and 36-55 represent more severely twisted bars.1
For each shape of test piece the shape of the plastic stress surface for the completely
plastic state is represented in an adjoining figure by contour lines (Figs. 36-41, 36-44,
36-47, 36-50, 36-53, and 36-56). The construction of these contour lines is very much
1 For details
regarding dimensions of the bars and applied torque moments ef. Z.
Ver. dent. Jng., p. 317, 1927.
FLOW LAYERS IN TWISTED BARS 525

facilitated by the condition that the surfaces in question consist entirely of either
planes or straight circular cones (cf. the wooden models and sand heaps on the pre
vious pages). Therefore, the contour lines consist entirely of straight lines and cir
cular arcs. In the schematic representations the penetration curves of these surfaces
are given. According to these assumptions, the elastic area in the cross section in the
completely plastic state must shrink until it coincides with the ridges (more exactly
the projections of the ridges) of the plastic stress surface which may be thought of as
erected over the cross section. It will be recognized by comparison of the etchings
with the contour figures that the middle lines of the white jagged-ridge areas in the

Fio. 36-57. Fig. 36-58. Fig. 36-59.


Flo. 36-57. Disturbance caused in elastic stress function after formation of first flow layer
in round bar. (After E. Reusa.)
Fig. 36-58. Stress surface corresponding to discontinuous yielding in wedge shaped flow
layers in round bar.
Fig. 36-59. Torque-twist diagram.

photographs agree quite well with the curves of the ridges of the plastic stress surface
shown in the sketches.1
36-6. Instability of Uniform Mode of Yielding in a Twisted Round Bar of Mild
Steel. The disturbance which must be caused in the linear elastic distribution
t = T«r/a of the shearing stresses in a twisted round bar of mild steel when the first
flow layers have formed in the interior of the bar (whose intersections with the plane
of a cross section appeared as black wedge-shaped narrow areas radiating inward in
Fig. 36-6) has been analyzed by E. Reuss1 in one of his noteworthy contribu
tions to the theory of plasticity in 1938. Reuss set himself the task of constructing
the elastic stress surface of a twisted round bar using the membrane analogy by assum
ing that the steel in one radial layer (along a radius in the circular cross section) has
been plasticized. Along this radial flow line he assumes that the metal in a very thin
1 The changes in microstructure of soft steel observed in the flow layers were utilized

by various investigators in further studies of the plastic deformations of this metal.


Cf. Turner, H., and J. Jevons, "The Detection of Strain in Mild Steels," Iron and
Steel Institute, Vol. 61, No. 1, 1925.
* Ueber Luders-Hartmann-sche Linien, Z. angew. Math. Mechanik, vol. 18, pp. 347-
357, 1938.
526 THEORY OF FLOW AND FRACTURE OF SOLIDS

layer appearing as a narrow band in the circular section has already reached the
"lower" yield point tj in simple shear while the shearing stresses t in certain regions
of the cross section have values t2 < t < ti, n being the "upper" yield point in sim
ple shear under which only elastic strains are produced. Thus he concedes the exist
ence of an unstable elastic equilibrium of the stresses in which the stresses t in a cer
tain portion of the bar overshoot elastically the lower yield value. Figure 36-57
reproduces this unstable equilibrium state in a round bar by means of the contour
lines F — const of the elastic stress function.
Although in these remarkable first attempts by Reuss the effect of the speed of
deformation on the yield stresses, which seems to have an important influence, was
not considered, they point to a fundamental problem of the instability of the uniform
elastic-plastic mode of yielding under torsion in a round bar as contrasted with a nonuni
form yielding in distinct plasticized wedge-shaped layers surrounded by elastic, regions.
If the volume under the elastic-plastic stress surface corresponding to the tentatively
guessed corrugated shape reproduced in the sketch (Fig. 36-58) by its contour lines
and to nonuniform yielding in radial layers with some definite average mean spacing
under a given constant rate of twisting should turn out to be smaller than the volume
of the stress function for uniform yielding in a growing ring-shaped area for the same
rate of head motion (Fig. 36-59), this might explain why nonuniform yielding in
twisted steel bars is frequently observable. The volume under the stress surface
being proportional to the torque M, this would physically explain why the bar yields
in discontinuous manner; the required torques are smaller than for uniform yielding,
and thus less mechanical work is needed to deform the bar.
The remarkable regularity disclosed in the flow patterns of twisted steel bars has
been emphasized by L. H. Donnell1 and its cause sought in an instability of the elastic-
plastic equilibrium of uniform yielding.2
1 "Plastic Flow as an Unstable Process," ASME meeting paper, December, 1941.
* See also Sec. 19-5 on instability of equilibriums causing certain types of fractures.
CHAPTER 37

PLANE STRAIN AND PLANE STRESS. THEORY OF


THE SURFACES OF SLIP
From my earliest experiments on the relation of electricity and
magnetism, I have had to think and speak of lines of magnetic power;
not merely in the points of quality and direction, but also in quantity.
Now it appears to me that these lines may be employed with great
advantage to represent the nature, conditions, direction and com
parative amount of the magnetic forces; and that in many cases they
have, to the physical reasoner at least, a superiority over that method
which represents the forces as concentrated in centers of action.
Faraday, Michael, "Experimental Researches in Electricity,"
Vol. 3, On lines of magnetic force, pp. 328, 329, 1855.

The tests on extrusion of certain ductile metals through orifices by


Tresca around 1868 led him to the conclusion that stressed metals start
to deform permanently when the maximum shearing stress reaches a
limiting constant value. Aiming at a quantitative interpretation of
these flow tests the great French engineer Barr6 de Saint Venant1 in 1871
advanced the first mathematical theory for the slow flow of solid bodies
in developing the theory of the two-dimensional or plane problem of flow
in a plastic material which is essentially the one described in the following
section 'for an ideally plastic substance.
37-1. Plane Strain. «« = 0. Suppose that in an extended body in
which a plastic state prevails the components of stress and of small strain
and the small components of displacement are functions of the rectangular
coordinates x and y only but do not depend on the coordinate z. Assum
ing the strain u = 0 in this case the components of stress tvi,tix and of
strain 7„i,y« vanish and the equilibrium of the stress components vx,vv,Txi
requires that
1 Compt. rend., vol. 70, pp. 368, 473, 1870; vol. 73, pp. 86, 1098, 1181, 1871; for a
critical review see also Todhunter and Pearson, "History of Elasticity and Strength
of Materials," Cambridge, 1886, and a paper by Marcel Brillouin, Ann. Phys.;
vol. 13, pp. 75-113, 1920. Saint Venant did not express the equations for the most
general case of three-dimensional plastic flow, stating that little hope could be advanced
for finding solutions of these equations. M. Levy (Compt. rend., vol. 70, p. 1323,
1870) proposed a set of equations for three-dimensional flow.
527
528 THEORY OF FLOW AND FRACTURE OF SOLIDS

dox , chjy
0, (37-1)
dx dy dy dx

These two equations are satisfied if oz,oy,Txv are taken equal to the partial
derivatives of a stress function F(x,y):

dW dW d2F
(37-2)
dy2 dx2 dx dy

A third equation for determina


tion of the unknowns o^Oy^^ is pre
scribed through the condition of
plasticity of the material. In an
ideally plastic substance, in which
T„ot = const, according to Eq. (27-9),

(o-x
- ovy + (o-y
— o-zy- + {o-2
- <rx)s

+ 6tV = 2<702 (37-3)

(co yield stress in tension). From


the third of the stress-strain rela
tions valid for infinitesimal strains:
<t>

6, =
\*x

'
2~7
a, + ox\
/

=
2—j
i>

e„

'
[cry (37-4)
, + Oy\
/

Ox
<t>

u =
\o, 2—
)'

Fio. 37-1. Le Comte Barre de Saint "Txy —


2(^rXB
Vcnant (.Courtesy of Pierre
,

(1797-1886).
Gaujat, Institid de France, Paris.)
one sees that, the normal strain
if

tt is taken equal to zero,

d'F d*F
Ox + 0-y dy2 dx2

more general case of plane strain characterized by assuming «, = = const.


1
A

is

Since the flow function = 7oct/3roct = 7ort/v/2°'o >n Eq. (37-3) for
$

V ca-o AF
2

+
<t>

7o«
2
2

to he taken. In this ease the equation corresponding to Eq. (37-6) above depends
is

also on the flow function and becomes good deal more involved than (37-6).
a
PLANE STRAIN AND STRESS 529

After substituting this value of a, in Eqs. (37-3)

(g,~g")2
+ r„2 =
£ = k* (37-5)

where the constant A; = <ro/y/% for an ideally plastic sub


is obtained,
stance. The square root of the left side of Eq. (37-5) represents the
maximum shearing stress 7W in a state of plane stress having the com
ponents at,,ay,,rsg. According to the maximum shearing stress theory which
was postulated by Saint Venant the same condition of plasticity holds
as in an ideally plastic substance under a
state of plane plastic strain e* = 0 but the
constant in Eq. (37-5) has to be taken equal
to k = <to/2.
A few analytical expressions for the three
components of stress a,,ay,r,./ and for

<T, =
id + <ru)/2

satisfying the two equilibrium equations FlQ 37-2


(37-1) in conjunction with the plasticity con
dition (37-5) will be considered below. Solutions of the plane problem
may also be found by substituting the three expressions in Eqs. (37-2)
for the stresses in Eq. (37-5), which leads to a partial differential equa
tion of second order and of second degree for the stress function

(dW _ dW\ ( d*F V = 4A;2


(37-6)

and by finding integrals of Eq. (37-6). In either of the two ways plane
states of strain in plastic bodies have been found satisfying certain
boundary conditions for the stresses along specific boundary curves
f(x,y) = 0.
Ifone desires to determine also the components of displacement in the
interior ofa body in a state of plane strain or in the case when the dis
placements are prescribed in the surface of a body or in parts thereof in
addition to the equations containing the components of 'stress, two further
relations must be considered. Let u and v be the small components of
displacement of a point x,y (assuming that w = 0) so that the plastic
strains:


du dv _ du dv
(37-7)
530 THEORY OF FLOW AND FRACTURE OF SOLIDS

<f>
After eliminating from Eqs. (37-4) using (37-7) the first relation con
necting stress and plastic strain obtained,

is
<ry~ *x _ (dv/dy) - (du/dx) Q7S)
2Tlv (du/dy) + (dv/dx)

to which may be added the condition of incompressibility,

£+5-°- <*-»'

Equations (37-8) and (37-9) determine u and the stresses a^v^r^

if
v
are already known. Equation (37-8) indicates that the principal direc
tions of stress and strain coincide.
The two components of stress <rz and <ry can be eliminated from the
.two equations (37-1) and the plasticity condition (37-5) by partially
differentiating the first of the equilibrium equations with respect to

y
and the second with respect to x. After subtracting one from the other
equation one obtains

dx dy
(o-," — "
<r„)
=
v

dx* dy*

and after substituting here for the stress difference its value az — <rv

taken from Eqs. (37-5) partial differential equation for the shearing
a

stress obtained
is

which has furnished some useful solutions.

Saint Venant made use of Eqs. (37-7) to (37-9) but he understood under u and
*

»
the components of the velocity vector and under «»,...
the components of the rates
of strain «, Instead of considering a state of equilibrium under steady slow
flow he expressed the equations of motion of a plastic mass as follows:
du du\ dat dr„ v
(du

J*
p\dtT T
U*Z + V*\ _«& T £»_ v
+

dy)
I[

dz dy dx
'

in the Eulerian form of the equations of hydrodynamics density, X,Y components


(p

of a body force). For steady slow flow Eqs. (37-4) and (37-7) to (37-9) may as well
refer to the components of small displacements u, and to state of equilibrium. It
t>

well understood that Eqs. (37-4) and (37-7) to (37-9) are valid expressions of stress-
is

plastic-strain relations provided that the strains and displacements are infinitesimal,
the principal directions of stress do not rotate, and all components of stress stay at
their given values corresponding to the limit of plasticity without undergoing any
further changes when the strains increase by infinitesimal amounts.
PLANE STRAIN AND STRESS 531

A group of exact functions describing plane plastic states of strain with


interesting applications were discovered using polar coordinates r,y>
instead of rectangular coordinates. If the normal components of stress
in radial and tangential direction are denoted by <r> and at and the shear
ing stress by Tr<, when r and <p are used as
independent variables, one has instead of i v«#

Eqs. (37-1) and (37-5) for the determination


of <rr,<rt,Trt the three equations

ST
~ <rt +
V ~
'

(37-11)
d<rt , d(rrr<) .

(<rr
- <n)2 + W .
(37-12)
Fio. 37-3.

Eliminating <r> and <r, from Eqs. (37-11), using Eq. (37-12) similarly
again a differential equation for the shearing stress rr, = t may be
obtained :

If a stress function F(r,<p) is introduced, by letting <rr, <r,, and rrt be equal to

one obtains for the determination of F the equation

These methods of finding exact functions for the components of stress


satisfying one or the other group of the preceding equations may profit
ably be combined with an analysis of the geometric patterns of the lines
of slip in these states of plane plastic strain. Under the latter we shall
understand the two systems of plane curves in which the cylindrical
surfaces of slip normal to the x,y plane bisecting the two principal
planes
intersect the
of stress at a point (x,y) which are normal to the x,y plane
latter plane. We shall see in Sec. 37-7 that the orthogonal patterns
of plane strain
of the curves of slip corresponding to the plastic states
have some remarkable geometric properties.
532 THEORY OF FLOW AND FRACTURE OF SOLIDS

<T,,<ry,T,y of a state of plane stress at a point


as the three components
of the x,y plane according to Eqs. (10-9a) can be expressed as follows:

<r, = —+5
<ri <r%
1
'
(<ri
— <rj)
o cos 2a
_
>
2 2

= <ri + <r% _ (ji —


<T2)
av cos 2a , (37-14)
! 2

(<rl <T2)
sin 2a ,

where <r\ and <r% are principal stresses, <r3 being taken equal to <rv a in
Eq. (37-14) is the angle which the algebraically larger of the two principal

,-33*. Aft
Fio. 37-4. Directions of principal Fio. 37-5. Lines of slip.
shear.

stresses a\ makes with the x axis. Since (<ri — <r2)/2 = equals the iw
maximum shearing stress in the x,y plane, at the plastic limit
— ,
(<ri <r2)
s = Kt = const ,
u

and if the arithmetic mean of the normal stresses is denoted by a, we have

<ri —
~~ <r%
1r — ' (37-15)
2 2~~

and Eqs. (37-14) may be written

a, = a + k cos 2a , <r„
= a — k cos 2a ,
= k sin 2a . (37-16)

In these three equations we can introduce a + 45 deg the angle /3


=
of the directions of the tangents to one of the two systems of the slip lines
by substituting, for the angle a of the directions of the major principal
stress <ri, a = /3 — 45 deg. The curves whose tangents are inclined under
these angles /3
= a + 45 deg may be called the first and the curves
inclined at the angles a — 45 deg the second system of slip lines. With
these stipulations:
PLANE STRAIN AND STRESS 533

-
a, = a + k sin 2/3 ,

ay = a k sin 20 , (37-17)
}
= — k cos 2/3 ,
and

tan 2/3 . (37-18)


2r.

If <r^<Ty,r,yare known functions of x and j/, the equations of the slip


lines y = /(x) of the first and second system can be found from the differ
ential equations

dy — cos ,
= tan /3
=
1

;
2/3
and
dx sin 20

- cotan — (3-M9)
("*

sin
/ I „\
dy , ^ 2/3
-ji = tan + = =
,

,
=

/3

1 -.
/3
I

dx — cos
\2 2/3

respectively, using Eq. (37-18). We shall see that this conveniently

y is
done after expressing both rectangular coordinates x and in parametric
form, as functions of the angle of inclination
/3.
37-2. Plastic Two Rough Parallel Plates.
Mass Pressed between
Using rectangular coordinates x,y suppose that the component of the
shearing stress t„ = f(y). From Eq. (37-10) we see that in this case
dh,y/dy2 = which satisfied we assume r,y = cy. Since according
is

if
0

to the plasticity condition Eq. (37-5) It^I Tn»i = there are two
^

straight parallel lines, for which we may assume = ±a, along which
y

the absolute value of the shearing stress Tx, reaches its maximum possible
value Taking = —k/a, = —ky/a will vary between the limits
k.

+fc and
— when varies between —a and +a. We conclude that for
k

the state of stress in which tw = —ky/a the two straight lines = ±a


y

natural plastic mass beyond which the mathe


boundaries
of

define the

matical solution ceases to exist. This strange attribute of plastic


a

distribution of plane strain in contrast to the behavior of the regular


is

solutions of problems in the theory of elasticity which may analytically


be extended beyond some arbitrarily assumed boundaries into an elastic
medium. It due to the character of the partial differential equation
is

for the stress function calling for certain properties of the solutions to
F

which further reference will be made in Sec. 37-8. Both stresses a, and
<ry may be calculated from the conditions of equilibrium (37-1) thus:
534 THEORY OF FLOW AND FRACTURE OF SOLIDS

The arbitrary functions fi(y), ft(x) must be so determined that the


condition of plasticity will be identically satisfied, thus:

— = ±2 — Tzy*
<Tx Oy \/k2

+ *+M-M*)- ±2kyjl- ■©"■


From this we have

- H

(J)'
yj\
My) = c ± 2k /,(«) =
a

;
Hence the stresses are determined by

— - - er

£
= 2k

\A
+

±
<r.
c

,
a

)
kx
=
H,

Oy (37-21)
>
C

T
- _M.
a

As these equations show, under the above assumptions,1 two different


solutions are obtained, according to whether in the expression for at
the positive or the negative sign taken. From the equations, a very
is

noteworthy peculiarity of these solutions may be discerned, namely, that


at both limits = ±a, both solutions give exactly the same stresses:
y

Txy =
-7 =
kx
+ — •
i

+
K

ff„
C

u
,

We now determine the equations of the slip lines under both stress conditions.
The proper sign to use in the right-hand member of the equation for a, in (37-21 may
)

be determined as follows, use being made of variable and Eqs. (37-17). According
0

to the latter we have

TzV
= —k cos 2/3 a, —
<r„
= 2k sin 20 (37-22)
,

Moreover, according to (37-21)

- - -
^

riv = and a, = ±2k \1 •


^

<r„ (37-23)

Therefore cos 2/3 = y/a and the sign of the radical must be chosen positive or nega
tive, according to whether the variable taken between < < r/2 or between
is

0
0

l/2 < < -K.


0

These solutions and the statement relative to the slip lines were first given by
1

L. Pbandtl, see Z. angew. Math. Mechanik, vol. 1923.


6,
PLANE STRAIN AND STRESS 535

The equations of the slip lines are determined from the differential equation in the
simplest way by using the variable 0 as a parameter.
Since,
y = a cos 20 ,
it follows that:

%-
dx
-2asin2/jp.
dx

The differential equation of the first family of slip lines

is,
according to (37-19)

:
^ - tan -2a

1^
= sin 20 ■

0
ax ox

That of the second family according to (37-19)


is

:
-r
ax
= — cotan 0
= —2a sin 20 -= •
ox

From these the following equations for the slip lines are determined

:
Firat family
X = -** + sin 2"] + COn8H (37-24)
1

= a cos 20
- sin 20]
v
\ / I

J
z = a [20 + const 1
Second family: — (37-25)
cos 20 .
y

If the quantity 20 replaced by a parameter (time) will be recognized that these


it
is

are equations of a family of cycloids, a being the radius of the rolling circle and the

%ty *k Txy *k
Fio. 37-6. Fio. 37-7.
Fios. 37-6 and 37-7. Slip lines and distribution of stress in a plastic mass pressed between
two rough parallel plates. Fig. 37-6 showing passive, and Fig. 37-7, active plastic state.

angle through which has rolled. The slip lines are represented by the two systems
it

of cycloids (Fig. 37-6) which cross one another at right angles. Each straight line
= ±a, along which the shearing stress reaches its maximum value and which
k,
y

rmt
we recognize as a natural limit of the plastic region, at the same time an envelope
is

of one series of slip lines. If the angle lies between and x/2 the first solution
0
0

results in a position of the slip lines according to Fig. 37-6, while r/2 < < t the
if

second solution gives the position of the slip lines according to Fig. 37-7.

Under conditions of the kind represented in Figs. 37-6 and 37-7,


a

soft, pasty, plastic mass squeezed between two wide parallel plates of
is

hard material, provided that the material may flow only toward one side.
To produce the stress distribution the solid plates must be brought
I

together. Under distribution II, however, the horizontal compressive


536 THEORY OF FLOW AND FRACTURE OF SOLIDS

stresses are the driving forces; under their action the horizontal plates
are forced apart. Since the limiting conditions relative to stress along
the horizontal compression plates y = ±a are the same for both stress
distributions, one may also — to differentiate both stress distributions
from one another — say, that on the compression plate in the state of stress
I a passive compression acts, while in the state of stress an active one II
acts.
Relative to the behavior of a plastic mass in the neighborhood of a
rigid rough surface we may therefore differentiate between two states:
State I
may be called the passive plastic and II
the active plastic state.1
The active and the passive flow may be visually represented by the cor
responding systems of slip lines. One may think of the state of stress I
(which corresponds to a positive sign of the radical or the region where the
variable /3 is between 0 and x/2) as represented by its slip lines in a plane
I, while the state of stress II (corresponding to the negative sign and
x/2 < /9 < represented in second plane II. These two branches
is
ir)

or regions of the solution may be thought of as coinciding along the


"branch lines" = +a to a single function. Plane contains the slip
y

I
lines of the passive and plane II
those of the active flow. The envelopes
of the slip lines are two "branch lines" of the solution along which two
states of stress,distinct in physical space, bound each other. We may
finally say: Solution Eq. (37-21) represents the plane flow of a perfectly
plastic mass, which has two parallel lines as "branch lines."

The relations near the free edges of a plastic mass compressed between two rigid
plates cannot be determined by the above equilibrium conditions. In the net of slip
lines, in the neighborhood of the free boundaries of the mass, new regions must be
considered in which the stress condition cannot be expressed by the Eqs. (37-21). It
difficult to predict how a continuation of the slip lines would appear, as long as the
is

other mechanical properties of the mass (its ultimate strength in pure shear and ten
sion) are not considered, since these determine the form of the slip lines at the edges
and free surfaces. On the basis of observations on distortion and slip, one would
expect that the slip lines in the neighborhood of free surface would tend to converge
a

toward an edge in the form of rays (cf. the photograph of the slip lines in plane com
a

pression test, Sec. 20-2, Fig. 20-20.) In this connection Prandtl has proposed possible
forms for the continuation of the slip lines.2 If the plastic flow may be considered as
two-dimensional, but the plastic mass permitted to flow away in both directions
is

between compression plates, there arises in the middle of the layer (in the axis of the
test piece) an area (Fig. 37-8) in which the shearing stresses (the friction) along the

This method of designation corresponds to the active and passive earth pressure
1

in the theory of earth pressure. In this respect there a near analogy between the
is

compressive action of soft plastic mass on a rough rigid plate and the pressure of
a

loose earth on a firm wall (supporting wall). Cf. the author's report on Plasticity in
"Handbuch der Physik," vol. Verlag Julius Springer, Berlin, 1928.
6,

See also Sec. 37-8 below.


2
PLANE STRAIN AND STRESS 537

compression plates do not reach their maximum possible value. It is only necessary
to consider that in the axis of the test piece the friction of the plates must be zero on
the basis* of symmetry conditions.
COMPRESSION
Hence, in this middle part of the mass,
the limits of plasticity will not be I I
reached. This region is shown dotted
in Fig. 37-8.
It should finally be noted that the
Eqs. (37-21) offer the possibility of
accurately determining the displace Fio. 37-8. Lines of slip for symmetrical case
of plane flow of a plastic mass pressed between
ments of the material in the plastic
two rough parallel plates.
layers. A system of components of
displacements u and v compatible with the system of slip lines of Eqs. (37-24) and
(37-25) can be found. The displacement component u parallel to the compression
plates is given by the ordinates of half of an ellipse, with respect to the variable y.

Fio. 37-11.
Figs. 37-9 to 37-11. Cross sections of plastic masses after compression between two rough
parallel plates showing the distorted forms of an originally rectangular network of straight
lines. Note the elliptic shape of markings in greater distance from line of symmetry in
Figs. 37-9 and 37-10 and points of inflection in markings near center line in Fig. 37-11.
(After W. Riedel.)

By squeezing out a highly plastic mass (wet clay or plastic plaster) compressed
between two parallel rigid plates W. Riedel has obtained deformations which support
538 THEORY OF FLOW AND FRACTURE OF SOLIDS

the conclusions arrived at above. In Fig. 37-9 the distorted form of an originally
rectangular network of straight lines can be recognized. The horizontal lines remained
parallel, but the vertical lines were transformed into curves very similar to halves of
ellipses as predicted by the theory. In Figs. 37-10 and 37-11 the central portion of a
plastic mass can be seen, flowing symmetrically to both sides. The latter case of
deformation corresponds to Fig. 37-8 of the slip lines and one can clearly see from
Figs. 37-10 and 37-11 that the ellipses are completely disturbed at the center of the
mass according to what must be expected from Fig. 37-8.

The slip lines in a slowly moving plastic mass should not be confused
with the streamlines of flow coinciding with the directions of the resultant
velocity vectors. In her careful studies of the properties of the surfaces
of slip Hilda Geiringer1 has also determined the streamlines for the case
of flow illustrated by Fig. 37-8. J
37-3. Radial Flow. In the case of radially symmetric plane flow the
radial and tangential normal stress a, and at are principal stresses and the
shearing stress rrl = 0. The first of the two equilibrium conditions
Eqs. (37-11) reduces to

£ (rar)
- ff«
= 0 (37-26)

and the plasticity condition to

or - a, = ±2fc. (37-27)

This leads to the well-known solution:

<jt = +2Jfc In -, a, = ±2k (l + In (37-28)


°)
(where a denotes an arbitrary length) which was derived in Chap. 30
for the plastic distribution of the stresses in a thick-walled cylinder.
Tfie slip lines are two orthogonal families of logarithmic spirals. These
spirals are commonly observed in plates or cylinders of low-carbon steels
whenever the yield stress is reached in them under radially symmetric
states of stress. W. Kriiger,* has observed such flow figures in the shape
of logarithmic spirals at one end of a thick-walled tube, which had been
1 "Fondemcnts mathematiques de la theorie des corps plastiques isotropes,"
Memorial des Sciences Mathematiques No. 86, Gautheir-Villars & Cie, Paris, 1937,
89 pp.
1 A series of flow tests exactly similar to the compression tests of W. Riedel has been
described by J. F. Nye, "Experiments on the Compression of a Body between Rough
Plates," Rept. No. 39/47 of the Physical Research Division of the British Ministry of
Supply, carried out at the Cavendish Laboratories, Cambridge, England, in 1947.
Nye used plates of a rolled tellurium-lead (0.05 per cent tellurium) deforming them
according to Figs. 37-9 to 37-1 1 above.
» Mitt. u. Forschungsarb., Ver. deut. Ing., vol.
87, 1910.
PLANE STRAIN AND STRESS 539

subjected to high internal pressure. Figure 37-12 shows similar markings


or flow figures on the polished surface of an iron piece obtained by forcing
a cylindrical punch into the test piece. From these markings we may
conclude that the forcing of a punch of cylindrical cross section into a
plastic metal, such as wrought iron,
tends to cause, at least in the thin
surface layer, a plastic flow of radial
symmetry having the radii and the
circles as the directions of principal
stress.1 These slip lines have no
envelope.

Fig. 37-12. Slip lines on surface of a steel Fio. 37-13. Slip lines
block produced by forcing a cylindrical consist of system of
punch into it. orthogonal logarithmic
spirals.

37-4. Circles as Envelopes of Slip Lines. We assume that the shear-


<t>.

ing stress rr< is a function of r only, not of denoted simply by t,


If
is
it

then Eq. (37-13) takes the simpler


form:

dr
+
3s-°- (37-29)
Fig. 37-15.
Fio. 37-14.
The solution of this equation Fio. The lines of slip are ortho
is

37-14.
gonal epi- and hypocycloids enveloping
two concentric circles.
= -j Fig. 37-15. The lines of slip are con
+

Ci (37-30)
T

centric circles and a pencil of rays.

To obtain the solution the boundary conditions


= = -Jfc
a
r

Slip lines in the form of logarithmic spirals may frequently be observed in the heads
1

of steam boilers subjected to plastic bending under an axially symmetrical stress field
(described, for example, by Korber and Siebel in Mitt. Kaiser-Wilhelm-Inst.
Eisenforsch.). These slip lines are also frequently visible around punched rivet holes.
In the well-known Brinell hardness tests, the metal obviously stressed in similar
is

way in ring-shaped area of the surface layer. Similar flow figures may be often
a

observed around the Brinell indentations while testing iron for hardness.

-
540 THEORY OF FLOW AND FRACTURE OF SOLIDS

and
r = b , t = +k
are used. Thus a ring-shaped plastic area is obtained, bounded by the
two concentric circles of radii a and b. These circles are the envelopes
of the slip lines. Equations (37-11) and (37-12) may be integrated com
pletely to determine the stresses and the equations of the slip lines.1
37-6. Vortical Flow in a Plastic Mass. We content ourselves in this case with the
statement of the formula for the case that the integration constant in Eq. (37-30)
cj = 0. We may then put:
a'
= —k- (r 1 a) , (37-31)

Furthermore, we introduce an angle 0, at which the first series of slip lines cross the
radius vector. We then have, corresponding to Eq. (37-17), page 533,

a, — a, = 2k sin 2/3
(37-32)
t = —k cos 2/3

also

cos 2/3 = -r (37-33)

The condition of equilibrium is

da,
r -2k (37-34)

from which

<rr=—2A-/^V7^+f. (37-35)

After introducing the variable /3 as given by Eq. (37-33), the integration may be easily
carried out, and the stresses o>,<r,,T determined as functions of the parameter 0:

a, = —k In tan

a, = —k In tan
(37-36)

—k cos 2/3

Moreover, the equations of the slip lines may be determined thus.


The equations for the first and second family of slip lines are given in parametric
form:
<t>

= arc tanh — arc tan = — arc tanh — arc tan + c,


v

v
v

- a— (37-37)
+

v*
1

r' = o, r2 = aJ
i

2v 2v

At a large diameter from the center the slip lines approximate logarithmic spirals
r

making an angle of 45 deg with the radial direction.

Cf. "Handbuch der Physik," vol. art. on Plasticity, Verlag Julius Springer,
1

6,

Berlin, 1928. The slip lines are epi- and hypocycloids.


PLANE STRAIN AND STRESS 541

A type of vortical flow compatible with this system of stress and slip lines is given
by the following displacements:

■5(->R) (37-38)

In this u' is the radial and v' the tangential displacement along the circle r >= const.
At the inner circle of the plastic area, the mass flows at an angle of 45 deg across the
circle r = a; its vortical motion, however, quickly dies down at a larger distance from
the center of the circle.
37-6. Radial Distribution of Stress in a Wedge-shaped Plastic Space. If the
shearing stress 771 depends only on the angle 4>, the differential equation (37-13) may

<j>,
be immediately integrated once with respect to thus
giving

= T2 VW^S
^

+ 2ck. (37-39)

In this the constant of integration designated by 2ck.


is

For = we obtain:
0,
c

Tn = sin (ci +
T

2<t>).
k

(37-40)

The corresponding stresses given by this particular solu


tion constitute a homogeneous tension or compression.
The integral curves (37-40) have, however, as may be
Fio. 37-16. The ray -shaped
seen, two envelopes, namely, the straight lines rr( =» ±k. region with extraordinarily
These are two singular solutions of the differential large shearing displacements
equation (37-39), no constant of integration appearing. radiating from the edge of a
If Tri = + according to the condition of plasticity compression plate.
k,

a, must equal a,. The two equations of equilibrium then give

= at = const
+

2k<j>
T

o> (37-41)

The stress distribution represented by the following equations


or = <n = +2k<t> const t„ = ±k
+

(37-42)
,

= const passing through the origin and the concentric


4>

has for its slip lines the lines


circles = const while the trajectories of principal stress are logarithmic, spirals.
r

This state of stress has a practical meaning, since may give the relations which
it

must exist in the sectors of severely distorted regions under compressive loading. If
a stiff plate pressed on a plastic mass, observations show that, on account of the
is

friction, ray-shaped areas extend from the edges. In these the material has suffered
extraordinarily large shearing displacements, as indicated by the sketch of Fig. 37-16.
(At this point reference may be made to the photographs of compressed test specimens
on page 335.) The points corresponding to the projections of the edges bounding the
compression surfaces of a test specimen of soft material, compressed between hard
plates, are singular points of the stress distribution. Wedge-shaped regions extend
from these straight edges into the inside of the specimen; the rates with which these
plastic strains change across the wedge-shaped region increase indefinitely the nearer
the point considered to the edge. The enormous rates, which must arise here, make
is

obvious why the sliding fractures in compression tests usually start along these
it

regions (pyramids and cones produced by fracture of brittle materials in compression


tests).
542 THEORY OF FLOW AND FRACTURE OF SOLIDS

It is noteworthy that the radial stress distribution of a plastic mass arises from a
stress function of the form F — fcr"£, which corresponds in the plane problem also to
a possible condition of equilibrium of a perfectly elastic material.1
If, however, in Eq. (37-39) c ^0, there results, as a more detailed investigation
shows,* a stress distribution in the form of a wedge-shaped area with two straight
lines as envelopes of the slip lines. Also this distribution of stress may be completely
determined. Two examples of the net of the slip lines are shown in Figs. 37-17 and
37-18.

Fio. 37-19. System of slip lines and dis- Fio. 37-18. Case where
tribution of stress in a plastic mass angle or of Fig. 37-19 is
pressed between two inclined rigid plates. a = ir/2.

A particular case is that of the equilibrium of a plastic mass compressed between


two rigid and rough plates, whose planes are inclined at a small angle to each other:

(37-43)

1
Regarding applications of the radial stress distribution given by Eq. (37-42) for
determination of the breaking conditions in a compressed test specimen of brittle
material, cf. Z. Physik, 1924, p. 106.
2 Z.
Physik, p. 106, 1924.
PLANE STRAIN AND STRESS 543

The compressive stress varies along the plates as the logarithm of the distance r
from the point of the wedge. In this a is equal to half the angle of opening of the
wedge (Fig. 37-19). The slip lines are given by the following equations:

In — — y/a* —
<t>'+ o arc sin — i
i'i a
(37-44)
In — — y/a* — —
<t>* a arc sin — ■
<•■; a

The above expressions, if the angle a is made infinitely small, reduce to those given
in the Eqs. (37-21), which apply to the case of a mass compressed between two parallel
plates.

37-7. The Equations of Plane Flow of an Ideally Plastic Substance


Expressed in Curvilinear Coordinates through the Slip Lines. The fre
quent appearance of lines of slip on deformed bodies justifies the raising
of the question whether from a mathematical point of view the surfaces
of slip accompanying the plastic distortion of materials might not offer
new advantages for expressing the equations of flow by means of a system
of these "natural" curvilinear coordinates. L. Prandtl, F. Koetter,
H. Reissner, W. Hartmann, and others have extended their application
to the equilibriums of several more general types of def ormable materials,
including loose granular substances (sand), in which two isogonal systems
of the surfaces of slip may be defined which are no longer orthogonal,
cases to which reference will be made in Vol. 2. Assuming an ideally
plastic substance, the two equilibrium equations

£' +
lf-°. £ + £"• ^-45)

and Eqs. (37-17) of Sec. 37-1 :

az = a + k sin 2/3 ,
try — a — k sin 2/3 , (37-46)
= — k cos
Txv 2/3 , ,

satisfying the condition of plasticity Eq. (37-5), let us consider the mean
stress a = (<rz + <r„)/2 and the angle of inclination /3 of the first system of
lines of slip as the dependent variables and, as Hencky1 first proposed two
parameters Ui and ut as the independent variables in the equations

<p(z,V) = Mi = const i>{x,y) = «2 = const


, (37-47)

through which the first and second system of lines of slip may be expressed.
After substituting the expressions Eq. (37-46) for ax,ay,Tzy in Eqs. (37-45)
1 Z. angew. Math. Mechanik, vol. 3, p. 241, 1923.
544 THEORY OF FLOW AND FRACTURE OF SOLIDS

we obtain

(37-48)

and after multiplying the first of these equations by cos 8 (and by — sin

f),
respectively), the second by sin (and by cos respectively) and adding:

8,
8
. if
')-.
1R

d<r
COS
+. r— cos + — sin
8

8
dx dy
(:

)-,(.
(37-49)
da .
/

cos
sin + fcoSll :)-„.

+
d/sinB
3
8

dy dx dy

The four parenthetical expressions in (37-49) are the derivatives of the


functions a and in the direction of the tangents to the slip lines. If «j
8

denotes the arc lengths measured along the lines of the first and s2 along
those of the second system as defined by Eqs. (37-47), Eqs. (37-49) may
be written in shorter form as follows:

+ = 4—{*- =
(a

2kB) 2k8) (37-50)


0

0
,

OSi 0S2

The arc elements ds\ and rfs2 may be expressed by

dsi = U2(u\,U2)du2 ds2 = U\(ui,u2)dui (37-51)


,

and Eqs. (37-50) may be integrated with the result:

f,
4+" /l(Wl) =
;

- =
/i(us) =
(37-52)
8

v%
,

2k

where vi = /i(mi) and y2 = fi(u2) represent two arbitrary functions of the


variables Ui and ui, respectively. Vi and t>2 are constant on a slip line
of their respective family.1 Therefore we see that the mean stress a
and angle are generally
8

= Vl = Vi — v2,
+

(37-53)
8

Vi>
2k

The equations of the slip lines may be given also by means of the functions:

>P*(x,y) = »i = const 4>*(x,y) = e2 = const . (37-52a)


,
PLANE STRAIN AND STRESS 545

a form of the solution of the problem of plane plastic strain due to


R. v. Mises.1

Equations (37-52) and (37-53) make it possible to deduce some interesting geometric
properties of plane fields of slip:
1. The mean stress increases or decreases along a slip line apart from a constant pro
portionally with the angle the tangent to
the slip line makes with a fixed direction
as the point of tangency is advanced
along it.
2. Draw the two tangents to two slip
lines of one and the same system at the \V/ :CONST.
points at which these lines intersect a line _ „- „n a a constant,
,
. . , r io. 37-2,. is
of the other system and repeat this for a
number of lines of the latter system. The pairs of such tangents are inclined to
each other at the same angle (a in Fig. 37-20) (Hencky).
3. Construct the slip lines using Eqs. (37-52a) for the parameters such as V\,
Vi + <>o,t>i + 2vo, . . . and »2, »2 + Do, t>j + 2w0, . . . , etc. The diagonal curves of
the resulting pattern of orthogonal lines are the isobars a = const (curves of constant
mean stress) and isoclines 0 = const (curves connecting points at which ft has the same
value) of the corresponding plane field of plastic strain (v. Mises).
4. Measure the radii of curvature of the slip lines of one system (for example,
ri = const) at the points at which a line of the other system (i>i = const) intersects
them. It may be shown utilizing (2) that the radii of curvature increase or decrease
by the lengths of arc measured on the line of the other system (Hencky, Prandtl).
5. If a line of slip is a straight line the curves of the system normal to it are concen
tric circular arcs in the neighborhood of the line. This is a corollary to (4). The
x,y plane may be covered without gaps by such triangular segments of infinitesimal
width in a continuous way. We see that systems of slip representing a class of exact
plastic states of plane strain exist in which the curves of one family are straight lines.
If the latter do not intersect in one single point they must have a common enveloping
curve. Let f(x,y) = 0 be the equation of an arbitrary plane curve. The tangents of
the curve f(x,y) = 0 and their normal trajectories, i.e., the tangents and evolvents of
a given evolute f(x,y) — 0 thus represent the lines of slip in a state of plane plastic
strain. Such orthogonal patterns of straight lines and their equidistant normal
trajectories have been discussed in Chap. 35 on plastic torsion of a prismatic bar.
They represented there the contour lines and the straight lines of maximum slope of
the plastic stress function, a surface of constant maximum slope in their projections
on a base plane. We therefore see that the trajectories of maximum shearing stress
(the curves whose tangents indicate the directions in which the resultant shearing

1 Bemerkungen zur Formulierung des mathematischen Problems der Plastizitats-


theorie, Z. angew. Math. Mechanik, vol. 5, p. 147, 1925. Equations (37-53) show that
a and 0 satisfy the partial differential equations:

^
dVi dv2
=0,' ^T=
dVi
dVi
0 (37-53a)

as was noted by C. Caratheodory and Erhard Schmidt (Z. angew. Math. Mechanik,
vol. 3, p. 468, 1923) and by L. Foppl ("Drang und Zwang," Vol. 1, 2d ed., p. 57,
Munchen, 1924).
546 THEORY OF FLOW AND FRACTURE OF SOLIDS

stress in the cross sections of a twisted bar acts) and their straight normals in a case
of plastic torsion may also be interpreted as possible systems of slip of plane problems
of plastic strain. The radii in a system of concentric circles and the latter, for exam
ple, represent a pattern of slip lines, which is the projection of the lines of maximum
slope and of the contour lines of a circular cone representing the plastic stress surface
of a round bar twisted around its axis. In a case of plane flow, this is the example of a
"radial plastic distribution" of stress a, = at =• ±2k<p, rri = ±k treated in See. 37-6.
In the general case of straight slip lines enveloping an evolute /(x,y) = 0 according to
(1) and Eqs. (37-52), the mean stress a = 2k(c + P) must be constant on any of the
straight slip lines 0 = const and increases or decreases proportionally with their
angles of inclination 0. On the evolute those values of the mean stress a will be found
which act along the respective tangents. In generalizing a former statement (Sec.
37-2), we may say that any evolute as envelope of a family of straight slip lines represents
a branch line in which two domains of the solution come in contact with each other
(in the sense explained in Sec. 37-2). 1

Fig. 37-21. Fio. 37-22.


Flos. 37-21 and 37-22. Systems of lines of slip.

6 The trajectories of principal stress in fields of plane plastic strain, according to


M. F. Sadowsky,2 have the property that they may be traced in an equiareal system,
i.e., that the parameters defining this system of orthogonal curves may be chosen so
1 We may note that the solutions of plane problems of flow behave in this respect in
manner analogous to that of the domains of multiple-valued analytic functions of a
complex variable. Similarly, as, for example, the domain of positive values of a
square root of a function of the complex variable z = x + iy may be represented in
one sheet, the domain of the negative values of the square root in a setond sheet, both
sheets constituting a Riemann surface, something analogous occurs with the solutions
in these problems of plane plastic flow in which the partial differential equations
are of the second degree. Their solutions "branch off" into two domains along
those lines, which are the envelopes of one system of the slip lines. See author's
paper, Tiber die Gleit- und Verzweigungsflachen einiger Gleichgcwichtszustande
bildsamer Massen, Z. Physik, vol. 30, pp. 106-138, 1924.
S. Khristianovitch (Moscow), in a paper ("The Plane Problem of the Mathe
matical Theory of Plasticity," translated by Brown University, Providence, R.I.,
1946) published in 1936, investigated the behavior of the solutions in the vicinity of
enveloping curves of the slip lines. For the latter he proposed the term "lines of
rupture." The term "branch lines" suggested by the author seems more appropriate,
since these lines have little to do with a fracture whose phenomena require entirely
different concepts, in spite of the fact recorded by Khristianovitch that certain space
derivatives of the components of normal stress become infinite along these lines.
' Equiareal Pattern of Stress Trajectories in Plane Plastic Strain, J.
Applied
Mechanics, June, 1941. Also Research Pubs. Illinois Inst. Technol., Mathematics,
vol. 3, p. 144, 1943.
PLANE STRAIN AND STRESS 547

that even the finite areas enclosed between consecutive curves of each system become
equal to a given constant. Orthogonal systems of curves in general have not this
property.
37-8. The Characteristics of the Partial Differential Equations in the Theory of
the Plane Flow of an Ideally Plastic Substance Are the Lines of Slip. We are here
concerned with certain special types of partial differential equations. We have seen
in Sec. 37-7 that it was possible, through certain clever transformations of the inde
pendent and dependent variables, to express the partial or total differential equations
to which the problems lead, in simpler form by reducing their degree from the second
to the first degree. In order to show the fundamentally different behavior of the
solutions of certain classes of differential equations well known to mathematicians we
shall examine briefly three types of linear partial differential equations of the second
order of a variable z dependent on two rectangular coordinates x and y, namely,

,. ,. d,z ,. .„_ ,A.


toi-ajT'"^'
d*z . d2z . d2z d2z , ,
& + 9?-t<**), Bx-Yy
"f(x'v)- (37-54)

We encountered an example for the first equation — + = Az = const in the


^-^
theory of torsion of an elastic bar, and examples for the second and third types may
be treated below. The first equation may define the equilibrium position of a thin
membrane uniformly stretched in its plane and carrying a transversal load distributed
according to a function f(x,y). Examples of the equations Az = const and Az = 0
occurred in Chap. 35, where it was noted that (1) their solutions corresponding to
given boundary curves can be determined as soon as the values of z are prescribed
along the boundary curve; (2) small irregularities in the boundary curve cause dis
turbances in the surfaces z which die down very rapidly around the former ones;
(3) the sharp concentration of stress in the interior of a twisted bar due to a small
cylindrical cavity in an elastic medium is very localized, but (4) in contrast to the
effects of a hole in an elastic medium and in a homogeneous field of stress at some dis
tance from the cavity as soon as plastic distortions have developed around the hole
the plastic regions extend far into the body.
a. The Field of the Displacements under a State of Uniform Stress. To examine an
example for the second type of Eqs. (37-54), suppose that we wish to determine the
components of small displacement u, and uy in a prismatic body subjected to a state
of simple tension or compression when the stress has just reached the limit of plasticity
while the body is not permitted to deform in the z direction. Neglecting the elastic
parts of the strains, in those portions of the body where the material has yielded, in
an incompressible substance the plastic strains must satisfy the equations

du, Oil, _ n
du,
y'"
du, du,
= n
+11 =0' "=17 =0'
= = + 0 - (37"55)
1^ ~dT

In the still unyielded portions of the body we assume that the components of strain
vanish. The third of Eqs. (37-55) states that since the x an y axes are principal
directions of stress and strain, the unit shear y,y must vanish for these axes. The
first of Eqs. (37-55) is satisfied if we let

If u, and uy are interpreted as the components of the small velocity vector of flow, <p
548 THEORY OF FLOW AND FRACTURE OF SOLIDS

represents the stream function. From the third of Eqs. (37-55) it follows that \p must
satisfy the partial differential equation

dT>
= (3"5°
dJ>

or an equation belonging to the second class of Eqs. (37-54) [assuming f(x,y) = 0].
Equation (37-57) is well known in mechanics. The deflections w of a vibrating string
stretched along the x axis satisfy an equation of similar form:

dhv 1 dhv
(3,"58)
W

~c'
ta~>

where the constant c2 = a /n, a the tension stress in the string and n the mass of unit

is
length, the time. Equation (37-58) the same as Eq. (37-57) we let = w,

if
is

^
t

.„ ct. We may, for example, think of plotting the integral w representing successive
y

configurations of an infinite string at the times as the function = w in the x,y plane

y ^
i
satisfying Eq. (37-57) by plotting the values of along the lines = ct = const. We

xfi
note that the motion of the string determined by two initial conditions, for example,
is

=0, w = f(x) and = g(x). One -

0,
dw/dt

t
t

conveniently changes the variables x,y — ct to


two new variables m and n by means of the
transformations

-—«-« ..—»+« o^9)


V2

'
V2
'
In terms of x and keeping the new parameters
y

m and n constant, the equations

+ y_m, ^L+l-n
5

(37-60)
Fio. 37-23. V2 V2
represent two systems of parallel straight lines inclined at the angles 45 and 135 deg
with respect to the x axis. Since

dj, = tydm dj, dn dA J_ df\


/

fd£
1-

— =
'

dm dx dndx y/2 \dm dn) y/2 \dm


'
dn)
(

dx dy

after once more taking partial derivatives, one finds that Eq. (37-57) [an example of
the second type of Eqs. (37-54)] transformed into the partial differential equation
is

''

~ 11
(37-62)
dm dn

[an example of the third type of Eq. (37-54)] having the general solution

- Mm) + f,(n) =
(~lXy)
/,

+
h

(^^) C37"63)
*

where and are two arbitrary functions. If the string allowed to carry on
/„
J\

is

a motion from an initially deflected rest position, shown in the uppermost curve
= w = f(x) in Fig. 37-24, the transversal velocity of point of the string at the
<l/

time — zero. Since = ct, = w from Eq. (37-63) after taking partial deriva
is

ip
0

y
t
PLANE STRAIN AND STRESS 549

tives with respect to y, we see that the condition y = 0, dw/dt = 0, or dtp/dy = 0 is


satisfied if fi'(x/y/2) = —fi'(—x/->/2). Furthermore, we have for y = 0,

/.(x/V2) +M—X/V2) -/(x).


We may now as well write }if(x + y) instead of f\[(x + y)/\/2] and — y)

instead of /2[(— £ + y)/\/2] and see that the sum of the functions Vif(x + y) and
— satisfies both boundary conditions: y = 0, ^ = /(i), and d^/dy = 0.
Hence, the deflections of the string at time t = y/c according to d'Alembert's solution

-*- + —

y)
+
J/)
u> (37-64)

willbe found by superposing these two functions of identical form [(these are two
waves displaced in their respective positions in opposite directions having one-half

Flo. 37-24. Deflections of string.

the amplitude of the original deflection w = f(x) at time = as indicated in the


0]
t

lower curves in Fig. 37-24.


Returning to the determination of the displacement components u,,u, compatible
with a uniform stress, let us assume that this uniform compression in the direction
is
a

of the axis, just at the yield point of the material. By virtue of Eqs. (37-60) to
y

(37-62), since the most general function satisfying Eq. (37-57) expressed by the
is

sum of two arbitrary functions, say F(m) and G(n), we have for the stream function

= F(m) + G(n) =
O
F

(37-65)
(IZ^M)
*

and the displacement components will be

«.-g-^=F(m) +(?'(«)],
(37-66)
u, - -*t - -L[—F'(m) +G'(n)]
dx V2
(the primes denote derivatives with respect to the arguments) or we should intro
if

duce the components of displacement um and un in the directions parallel to the


550 THEORY OF FLOW AND FRACTURE OF SOLIDS

straight axes denoted by m and n in Fig. 37-23, u„ and un will be given by

V2G'(n),
(36-67)
\/2F'(m).

These last two equations obviously define two states of simple shear of variable
intensity, since in each state the displacement is a function of the coordinate normal
to its direction. The functions G'(n) and F'(m) are determined through the displace
ments prescribed on the edge y = 0. The analogy1 between the representation of the
deflected shapes of an infinite string as a function of the coordinates x at the time* t and
the problem of the determination of the components of the displacements in a body stressed
in compression becomes evident. Similarly, as deflections of equal magnitude in the
successive snapshots of the deflected string represented in Fig. 37-24 in the x, y = ct
system of rectangular coordinates are propagated along two systems of inclined
straight lines known as the characteristics of the differential equation of the deflection
Eq. (37-58), the components of displacement u. and un of equal magnitude in the
compressed body will be found along two sets of parallel straight lines inclined at the
angles 45 and 135 deg with respect to the x axis. In these lines, the characteristics of
Eq. (37-57) x + y = y/2 m and —x + y = y/2 n, we recognize the familiar lines of slip
in a compressed body coinciding with the directions of maximum shearing strain.
We noted that the most general integral of the partial differential equations of the
second and of the third type of Eqs. (37-54) is expressed by the sum of two arbitrary
functions whose arguments are the parameters m and n of the characteristics of these
equations. This form of the general integral is particularly adapted and suitable for
solving those problems in which discontinuous solutions of these partial differential
equations must be considered. Under these we shall understand here primarily
cases in which one of the dependent variables (<//, w, u„ or u„, . . .) whose values are
prescribed along the boundary line y = 0 of the x,y plane vanishes in certain intervals
of the variable x and has finite continuous values in other intervals, although the
method of utilizing the characteristics is more general and may be applied even in
cases in which the function has finite discontinuities at certain values of x along the
line y = 0. A few examples of systems of parallel layers of slip in which the compo
nents of displacement in an extended or in a compressed body behave in the former
manner are illustrated in Figs. 37-25 to 37-27. Two such layers or several parallel
strips may cross each other in which the material is distorted by simple shear of vari
able or uniform intensity and these layers may be separated by domains of the body
which remain undistorted and are merely displaced in the x,y plane as rigid bodies.*
1 Zur Theorie plastischer Zustande, Proc. 3d Intern. Congr. Applied Mechanics,
Stockholm, 1930.
* Incidentally, it may be noted that among the systems of displacement satisfying
Eq. (37-57) states of strain may occur in which only one layer of slip will form in the
direction of one of the two characteristics m = const or n = const. This is the case,
for example, if one portion of the body adjoining the edge is fixed in space, while the
other part remaining rigid beyond the slip layer is allowed to move as a rigid body
according to the motion prescribed through the shear in the layer. Thus we see that
slip along a single family of characteristics is also a possible case of distortion and that
the requirement of the Mohr theory claiming the formation of two symmetrical systems
PLANE STRAIN AND STRESS 551

In summing up, we see that some properties of the three simple types of linear par
tial differential equations of the second order introduced in Eqs. (37-54) may be dis
tinguished. A solution of the first type of these equations
an a*x .. .

lying within a closed curve (which


as a rule is to be defined for an area in the x,y plane
may extend to infinity) and is determined through the values of either z or of dz/dn
along the given curve. In contrast to the behavior of the solutions of the first type

Fig. 37-25. Fig. 37-26. Fig. 37-27. Layers of


Figs. 37-25 and 37-2,. Two intersecting layers of shear in compressed body,
slip. (Left) Tension. (Right) Compression.

of Eqs. (37-54) a closed area need not be considered for determining a solution of the
second or of the equivalent third type of Eqs. (37-54). It does not suffice to prescribe
one boundary condition, but it is necessary that two conditions be prescribed along

of surfaces of slip is not a necessary attribute in their laws of formation. It may be


<l>,

shown that any of the quantities u,, u„, u, «„, 7m. satisfies an equation of the type
dh///dm dn " having the straight lines Eq. (37-59) for characteristics. It should
0

be added that the functions F'(m), G'(n) in Eq. (37-67) must satisfy certain require
ments: Only monotonously increasing functions F', G' with their respective arguments
are permissible; otherwise the energy law would be violated within the plasticized
layers.
The preceding theory in which plastic layers were assumed embedded between
perfectly rigid regions of a body which were displaced without suffering any distortion
does not yet represent a complete way of describing the interesting phenomena of the
formation of the layers of slip because the elastic distortion of the material has been
neglected in the yielded and unyiclded portions of the body. It may be worth noting
that, in the case of more general plastic states of stress when the slip layers are curved,
the requirement of having certain regions displaced as rigid bodies between the wedge-
shaped plastic layers may not so easily be satisfied as for straight characteristics.
552 THEORY OF FLOW AND FRACTURE OF SOLIDS

the given portion of the boundary, e.g., the values of z and of dz/dn. The solution
may be completely denned inside two strips lying between the two pairs of characteris
tics which may be drawn through the end points of the given curve without a recourse
to further values of z and dz/dn beyond the given curve. The solution z satisfying
the second (or third) of Eqs. (37-54) will thus be determined in the triangular area in
which the two strips lying between the two pairs of characteristics intersect each other
on one side of the given boundary curve.
6. Characteristics of Partial Second Order Hyperbolic Differential Equation of Tuo
Independent Variables. In the theory of partial differential equations of a variable z
dependent on two variables x,y which are linear in the second order terms and whose
most general form may be assumed as follows,

, 32z . nD d*z . „dH ,( dz dz \

where A,B,C are functions of x,y and the function/ need not be linear in z, dz/dx, dz/dy
it is shown that the solutions of such an equation satisfying general boundary condi
tions may be constructed.1 Relative to Eq. (37-68) a problem stated by Cauchy may
be investigated: let a plane curve I in the x,y plane be given through g(x,y) = 0.
Along I erect given values of the variable z perpendicular to the x,y plane and draw
through the end points of these ordinates z straight lines inclined under the given
slopes dz/dn in the direction normal to curve I. These lines define a strip of a develop
able surface passing through a curve in space (whose projection on the x,y plane is
the given curve g(x,y) = 0). According to Cauchy, a surface z = z(x,y) satisfying
the differential equation (37-68) is sought which is tangent to the developable surface
along the given space curve. The conditions prescribed along curve I may be inter
preted as the "boundary conditions" imposed on the solution z of Eq. (37-68).
While the question whether a solution z may be found depends on the coefficients
A,B,C, cases may occur for which an infinitude of solutions satisfying the given bound
ary conditions along certain definite groups of curves may exist; moreover, for a
certain class of the general equation (37-68) boundary conditions in the form just
stated cannot be assumed, and for this class of differential equations either the ordi
nates z or the slopes dz/dn can be prescribed, but not both, along a given curve /.
After making use of the standard abbreviations:

rp =
dz
— ', g=T->
*
dz d*z
r = T~i ' s " d*z
IT T" '
„ d'z
1 = -ST!
,„_ „
(3"-C8a)
,
dx dy dx2 dx dy dy2

■See the theory of characteristics in Forsyth, A. R., "Theory of Differential


Equations," Part IV, Vol. VI, 596 pp. (Cf. Chap. 20, p. 388), Cambridge University
Press, London, 1906; Frank, P., and R. v. Mises, "Die Differential und Integral-
gleichungen," 2d ed., Vol. 1, pp. 681, 783, R. Oldenbourg, Milnchen and Berlin, 1930;
Sommekfeld, A.,
" Vorlesungen iiber Theoretischc Physik," Vol. VI, p. 36, Partielle
Differential Gleichungen, Dieterich'sche Verlagsbuchhandlung, Wiesbaden, 1947,
332 pp.
The characteristics of the partial differential equations of the two-dimensional
flow of a compressible fluid have been used extensively (after certain transformations
were carried out on these equations) by J. Ackeret, A. Busemann, T. v. Karman,
L. Prandtl, and others for solving important problems on the flow of compressible
fluids. In this connection cf. Liepmann, H. W., and A. E. Puckett, "Introduction
to Aerodynamics of a Compressible Fluid," John Wiley & Sons, Inc., New York,
1947, 262 pp.
PLANE STRAIN AND STRESS 553

denoting by a the angle of an element of are dm of the given curve g(x,y) = 0 with the
x axis and after recalling that Cauchy's boundary conditions prescribe the ordinates z
and the gradient (grad z) of the surface z(x,y) along curve I (vector grad z may be
resolved into its rectangular components p = dz/dx, q = dz/dy or its components
dz/dm, dz/dn) we note that p and q:

dz
rp = — = cos a —
dz

dz
sin a — >
dx dm dn
(37-68/<)
= sin a — + cos a ---
dz
— . dz . dz
= -
*q
dm dn

as well as the differentials dp and <fy:

Xv.tX-tt)
will be known quantities along curve i and that together with Eq. (37-68) or

Ar +2Bs + Ct =f (37-68d)

we have obtained three linear equations (37-68c , d) for the three unknown second deriv
atives r,s,t along /, from which, by eliminating two, e.g., the unknown s may be com
puted from the equation

[A (dy)'
- 2B dy dx + C(dx)2]s

(A dp dy + C dq dx + f dx dy) = 0 (37-68f)

and similarly r and t.1


If the determinant of Eqs. (37-68c, d)
A 2B C
D = dx dy 0 A (dy)* -2Bdydx + C(dx)* = 0 (37-68/)
0 dx dy\

vanishes, both brackets in (37-68e) being equal to zero, the values of r,s,t cannot be
computed. This results in a quadratic equation for the ratio of the differentials dy/dx.
Assuming that the discriminant B2 — AC > 0, through the condition (37-68/) at any
point of the x,y plane two real ratios dy/dx are defined. We see that in this case two
new families of curves in the x,y plane may be constructed, say

= t „• const = r, = const .
<p(x,y) , <p(x,y) (37-6%)

which will be tangent to these two directions and which are called the characteristics
of Eq. (37-68), since their form depends only on the coefficients A, B, and C of Eq.
(37-68). Suppose that we let our curve / coincide with one characteristic; we may,
for example, assume the distribution of the values of s along it arbitrarily and may
compute the corresponding distributions of r and f from Eqs. (37-68e, d). Thus an
infinitude of solutions z of Eq. (37-68) and integral surfaces z(x,y) satisfying two given
boundary conditions (z, dz/dn) along the chosen characteristic may be obtained. In
the theory of partial differential equations of this type (for which O=0 and the dis
criminant is positive) one proves that these surfaces z{x,y) having the same values of
1 If the expression in the first bracket of Eq. (37-68e), representing the determinant
D of Eqs. (37-68r , d) is not zero, this process may be repeated for the higher derivatives
of z, and it may be shown that a solution z of Eq. (37-68) exists satisfying the boundary
conditions and that the higher derivatives of z along curve I may be computed.
554 THEORY OF FLOW AND FRACTURE OF SOLIDS

x,y,z,p,q along the characteristic touch each other and furthermore that, if the parame
ters £ and >j of the characteristics are introduced as new curvilinear coordinates,
Eq. (37-88) is transformed to the normal form

d'z
= function (z , -? '
d£dv
.
-£ J
>{,17 (37-68* )

When the discriminant of Eq. (37-68/) is equal to zero the two systems of the charac
teristics coalesce in one family of curves, and when it is negative no real characteristics
exist. The three corresponding types of partial differential equations are known as
hyperbolic, parabolic, and elliptic.1
It is understood that a partial differential equation of the general form of Eq. (37-68)
may change its character in different regions of the coordinate field and may be hyper
bolic in a certain region and may become parabolic or elliptic in adjoining regions, etc.
c. The Two Systems of Orthogonal Lines of Slip in Plane Strain of an Ideally Plastic
Substance Are the Characteristics of the Differential Equations of Plane Plastic Strain.
The characteristics are straight lines under a state of homogeneous strain, but under
nonhomogeneous states of plane strain in an ideally plastic substance they are in
general two systems of orthogonal curves. The patterns of straight lines combined
with equidistant curved lines of slip and the various orthogonal systems of curved
slip lines which were determined in the examples mentioned in Sees. 37-2 to 37-6
show us that the characteristics of the two partial differential equations (37-53a), valid
for the mean stress a and for the angle of inclination fi of the directions of slip in nonhomo
geneous plane states of plastic strain in general are two systems of orthogonal curves
[Eqs. (37-52a)]:
v*(x,y) = Vl , +*(x,y) = v, . (37-68t)

According to H. v. Mises* the system of displacement components for plane plastic


strain, similarly as it was shown for a state of homogeneous stress for the components
ux,uv in rectangular coordinates, may be derived in the case of nonhomogeneous stress
from a stream function <j/. This function must satisfy the partial differential equation
of hyperbolic type

dv, ft, *
dVi dv2 dt'i R-i

again having tlie curved slip lines for its characteristics.3


Moreover, it may be seen, e.g., in a plastic body stressed under a state of homo
geneous plane strain under the principal stresses ax and <rv, the shearing stress rf, being
zero, that the stress function F satisfies the differential equation

d'F d*F =
., ±2k (37-69a)
dy* dx1

*
If, for example, / Eq. (37-62) in Sec. 37-8 is obtained, whose characteristics
= 0,
are two parallel systems of straight lines.
1 The first of the
three equations (37-54) is thus of elliptic, the second and third of
hyperbolic type.
1
Loc. cit.
*
Ri and Ri are the radii of curvature of the lines of slip. Applications of Eq. (37-69 i
among other examples were treated by It. Lohan (Dissertation, Berlin, 1935) dealing
with the plane problem of a plastic mass.
PLANE STIiAIN AND STRESS 555

the solution of which is

*> + ' (%y) + "


i^f)
-
<37-70>

Remembering that in an infinitesimal region around a point x,y in a variable stress


field the state of stress is homogeneous when singularities are excluded, we may con
clude from the preceding that it should be possible, as in the case of Eqs. (37-53a), to
"linearize" the partial differential equation (37-6) for the plastic stress function F by
expressing F as a function of the parameters Vi and t>2defining the orthogonal system of
curves coinciding with the lines of slip in the general case of plane plastic strain, i.e.,
by introducing as curvilinear coordinates the characteristics of Eg. (37-6).
In contrast to the behavior of the integrals of partial differential equations of
elliptic type (the first of Eqs. (37-54)] integrals of the hyperbolic equations having
different analytical expressions in two independent variables may be patched together

Fia. 37-28. Fio. 37-29.


Figs. 37-28 and 37-29. Plastic mass compressed between rough parallel plates. Patterns
of the characteristics (lines of slip). (After L. Prandtl.)

along characteristic curves in the x,y plane. Thus an integral function satisfying cer
tain prescribed boundary conditions along a curve may be extended beyond the
domain in which it is valid into an adjoining domain in which another integral func
tion may be used, and this process may be repeated in new domains without the
necessity of changing the values of the integrals in the former regions. Adjoining
domains having distinct integral functions or solutions of the differential equation
have a common characteristic curve as their boundary.
This may be illustrated qualitatively by two patterns of slip lines reproduced in
Figs. 37-28 and 37-29, due to L. Prandtl,1 showing slip lines radiating from the edge of
a body compressed between parallel rigid plates and in the interior. In the segments
having straight slip lines radiating from the edges of the compressed body (the points
A and B, Figs. 37-28 and 37-29) the state of stress is the one which was described in
Sec. 37-6, while in the region to the right of DFE (Fig. 37-28) the state of stress coin
cides with the solution discussed in Sec. 37-2 leading to the orthogonal systems of
cycloids as slip lines; within the triangles ABC a state of simple uniform compression
acts. Thus we see again that the characteristics represent mathematically strips of
infinitely small width along which two different integrals of a given equation may be
fitted together. They may also be described as those curves (or surfaces in space)

1 Spannungsverteilung in plastischen Koerpern, Proc. 1st Intern. Congr. Applied


Mechanics, Delft, Holland, 1925.
550 THEORY OF FLOW AND FRACTURE OF SOLIDS

along which local disturbances created at a point of the field are propagated through
the medium.1
37-9. Plastic States of Plane Stress. In thin disks the normal stress a, and the
shearing stress components rz, and t„, are equal to zero. The stress components
a,,a„ and t,„ in the plane of the disk satisfy the equations of equilibrium (37-1) and
the condition of plasticity Eq. (37-3) which now takes the form

<r*2
- <r*r, + <7„* + 3tx„2
= <r„* . (37-70a)

If the normal stresses are principal stresses <7i,<r2as the stresses a, and at in radial and
tangential direction in a rotationally symmetric case of stress (rrJ = 0), Eq. (37-70a)
reduces to
<Tlff2 -f" ff22 = 0a2

Cl2 (37-71 )

or to a plasticity ellipse. The latter case has been discussed in detail in Chap. 34.
W. Sokolowsky2 remarked, when discussing the general plastic state of plane stress,
that the equations of equilibrium (37-1) in conjunction with Eq. (37-70n) may be trans
formed in a manner analogous to the way in which the corresponding equations for
the state of plane strain were treated in Sec. 37-7. He found that under plane stress
cases can occur in which the differential equations are of hyperbolic or elliptic type in
distinct regions of one and the same disk, with the result that real characteristics were
found only in the zones corresponding to the former type of equation.*

37-10. Conclusions. When considering the multitude of phenomena


associated with plastic states of stress, with the formation of the surfaces
of slip, and with the criteria of rupture in solids described in previous
chapters, it appears that their problems were treated historically and
might still be discussed in perhaps three different ways worth being at
least briefly characterized. The earlier investigators approached some of
these problems by attempting to establish general criteria on which certain
types of the conditions of failure would depend. Coulomb, Rankine,
1 This becomes evident in the problems of hydrodynamics of compressible fluids in
which these surfaces have been studied and utilized particularly in the theory of super
sonic flow.
2 Two of his papers in Compt. rend. acad. sci. U.R.S.S., vol. 51, No.
3, pp. 175-1 7S;
No. 6, pp. 421-424, 1946 (published in English) ; also Equations of Plastic Equilibrium
for a State of Plane Stress, Prikladnaia Matemalikai, vol. 9, pp. 111-128, 1945, trans
lated by Brown University, Providence, It. I., 1946.
* Sokolowsky shows, for example
(see his book, "Theory of Plasticity," Chap. 10,
p. 214, Published in Russian in 1946), that the characteristics (lines of slip) around a
circular hole of radius r = a in an infinite disk yielding completely may consist of
certain curves which will terminate abruptly along a certain circle of radius r = b, if
the disk is stretched uniformly by the stresses a, = <r( = aa for large values of r but
is loaded by a pressure aT = — p in the hole r = a. Real characteristics are found
only within the ring-shaped region air id;
the characteristics intersect the inner
circle r = a at angles of 45 deg, but both curve systems become tangent to each other
along the circle r = 6 in which they terminate abruptly. In the region b < r < <*>
the equations become elliptic and no real characteristics exist. He discusses also a
number of interesting cases of the characteristics around an elliptic hole.
PLANE STRAIN AND STRESS 557

Duguet, Mohr, and others did not attempt to formulate the mathematical
problems as generally as they were treated later by St. Venant and by
the investigators during the second quarter of the twentieth century, who
began to utilize in their treatments the methods that had been developed
for similar purposes in the mathematical theory of elasticity and in the
mechanics of continuously distributed masses (hydrodynamics). It is
believed that, in spite of the incompleteness of the treatment of some of
these questions by the early originators of the theory of plasticity and of
fracture, considerable credit should be given to the work of the investi
gators of these problems during the major earlier portion of the nineteenth
century, and some of this work of engineers should have received more of
the justly deserved acknowledgment than it has found, particularly
among metallurgists and physicists. It is to be emphasized that the
concepts earlier announced of the limiting surface of yielding as repre
sented in a rectangular system of variables in which the principal stresses
<ri,<Ti,<t3 serve as coordinates and similarly, the concept of the Mohr enve

lope of the major stress circles in the <t,t plane of stress, if combined with
the conditions of equilibrium in the example of the two-dimensional prob
lems of plastic distortion and flow, contain the essential prerequisites
for the theory of the surfaces of slip representing the characteristics of
the corresponding differential equations. Mohr was certainly aware of
the fact that he had formulated rules for defining the orientation of the surfaces
of slip under very general properties of solid materials, among which he
included, for example, a theory of the surfaces of slip in a loose granulated
material. Whereas in some newer work in which authors have elabo
rated upon fracture hypotheses the impression has been left that the newer
hypotheses would offer fundamentally different or improved approaches,
e.g., in the discussion of theorientation of the planes of shear or of cleav
age relative to the directions of principal stress, from those previously
considered but recently ignored or believed to be less satisfactory; in some
instances perhaps the contrary is the truth.
As already mentioned, one may, following a suggestion of L. Prandtl,
rationally define a general class of plastic substances, basing it on a gen
eralized condition of plasticity in accordance with the thoughts of Mohr
and on the equation of an envelope of the major principal stress circles

<ri
- <r3
= /O, + <r3) (37-72)

assuming that <ri > <r* > <r3 and omitting to include in it the intermediate
principal stress <r2. If this equation is combined with the two conditions
of equilibrium of a state of plane plastic strain containing the components
of stress ir,,<ri(,T,S since
558 THEORY OF FLOW AND FRACTURE OF SOLIDS

three equations are obtained defining <sz,ay,Txu. Recalling the simple rule
established previously in Sec. 15-13/ for the construction of the angle
of slip in a point of the Mohr envelope, a number of interesting and
important conclusions might be drawn, say, in the simplest cases of
homogeneous distributions of stress ax = const, <rv = const, t^ = const
predicting, for example, the angles of slip, the angles of shear faults and
of jointing in large bodies of rocks, etc., without having recourse to a
discussion of the characteristics of the three aforementioned equations
under nonhomogeneous states of stress, whose angle for the latter states
will be found to vary from point to point. A clear case might be worked
out postulating an ordinary parabola as Mohr envelope, according to
A. Leon and C. Torre.1 This then would represent a first way for dealing
with general cases of permanent distortion including formation of surfaces
of weakness, in geologic applications, for example, of the faulting and
jointing in large masses of rocks or in engineering applications considering
brittle materials having a much higher compression than tensile strength,
predicting their strength under any other general state of stress in good
agreement with observable values and of the variations of the angles of
the surfaces of fractures with the states of stress, characteristic of such
substances.
A second method which has been proposed more recently for dealing with
similar cases under special and mathematically simpler conditions of flow
makes use of the calculus of variation, by expressing certain resultant
forces or the work dissipated in an ideally plastic substance and determin
ing the state of stress by making the efforts or the work an extremum
under additional conditions of constraint (Sadowsky, Hill, Bijlaard, and
others). It may be stimulating to extend some of these methods to
plastic substances of more general character such as the ones whose con
dition of flow considers a Mohr envelope touching the major principal
stress circles representing the limiting states of stress.
The third geometric method for dealing with these problems of plane
plastic strain is the one described in the preceding sections for the ideally
plastic substance which utilizes the mathematical simplifications of the
differential equations made possible if the parameters of the characteris
tics are introduced as coordinates. The method may be generalized to
include the cases when the condition of plasticity is expressed by a more
general Mohr envelope [Eq. (37-72)].

1 Loc. cit.
PLANE STRAIN AND STRESS 559

Further ways for attacking these problems might be based on an equa


tion of the surface of yielding

/i(«ri,c2,ffj) = 0 (37-73)

instead of on a plane Mohr enveloping curve [Eq. (37-72)] and by con


sidering plausible special forms of such as were proposed in Sec. 15-lSh,

it,
again leading to three approaches for obtaining the solutions entirely
analogous to the three types which were just described. Equation
(37-73) in conjunction with the equations of equilibrium may be treated
as well as Eq. (37-72) by means of the method of characteristics in
problems of plane strain and useful cases worked out under the more
general assumption of a condition of flow made dependent, for example, on
the mean stress a = (at
+
+

<rj <r3)/3.
AUTHOR INDEX
Boardman, H. C, 179
Boas, W., 43, 44, 51, 53, 184
Ackeret, J., 552 Bohuczewitz, 246
Adams, L. H., 34 Boker, R., 66, 238, 239, 241, 243
Allen, D. N., 467 Bollnow, O. F., 38
Amsler, 247, 255, 332 Boodberg, A., 267
Anderegg, F. O., 198 Born, M., 38, 40
Anderson, A. It., 78 Boussinesq, 151
Anderson, It. L., 82 Bragg, Sir Laurence;, 38, 61
Andrade, C, 50 Bragg, W. II., 38
Ansel, G., 53 Brand, L., 152
Anthes, 499 Brandtzaeg, A., 238
Arnold, E. E., 344 Bridgman, P. W., 30-35, 58, 81, 177, 271,
Aronofsky, J., 82, 319 273, 442
Auerbach, 400, 404 Brillouin, M., 405, 527
Brincll, 539
IS Brinkmann, H., 21
Brown, A. L., 266
Bach, C. von, 26, 297, 298, 300, 353, 358, Brown, R. L., 238
375 Brown, W. F., 261
Bader, W., 20, 512 Burgers, T. M., 16
Bailey, It. W., 46, 417 Busemann, A., 552
Baker, J. F., 375
Baker, T. C, 188
Baldwin, W. M., 428
Bardenheuer, 26 Carat heodory, C, 545
Barrett, C. 8., 52, 53, 56, 199 Carlson, R. K., 187
Barstow, F. E., 193 Carpenter, 55
Baud, R. V., 289, 291 Cassehnum, 21
Baumann, R., 26, 300, 353, 358, 375 Cauehy, A. L., 112, 552, 553
Becker, It., 62, 63 Chalmers, B., 67, 68
Belaiev, N. M., 404, 437 Cloog, M., 196
Beltrami, 209 Cloo-s, H., 223
Benedicks, C, 189 Coffin, L. F., 441, 455, 467
Berliner, 27 Coker, E. G., 503
Bernouilli, 37 Considere, 71
Betty, B. B., 53, 54, 67 Contorova, 195, 201
Biggert, F. C, 281 Cook, G., 297
Bijlaard, P. P., 322, 325, 326 Coulomb, C. A., 219, 220, 225, 406, 467,
Bingham, E. C, 15, 404, 420 556
Birkhoff, G., 36 Cozzone, F. P., 358
Bleakney, H. H., 194 Cretin, 517
Bleich, F., 370 Cromwell, F. J., 274
561
562 THEORY OF FLOW AND FRACTURE OF SOLIDS

Cunningham, I). M., 269 Forsyth, A. R., 552


Czochralski, J., 43, 44, 515 Frank, P., 92, 552
Frenkel, J., 178, 195
Friedman, H., 499, 521
Friedman, J. B., 185
Davidenkoff, N. N., 78, 80, 85, 190 Frocht, M. M., 394
Davis, Evan A., 20, 84, 180, 191, 213, 254, Fry, A.. 279, 284, 285, 340, 367, 512
256, 259, 260, 298, 310, 349, 412, 413,
442 G
Davis, Harmer E., 83, 85, 185, 265, 267,
270 Galin, L. A., 505
Dawance, G., 350 Gauss, 169
DeForest, A. V., 78 Geckeler, F. W., 375
Deutler, 21 Gehler, W., 375
Dollins, C. W., 53, 54, 67 Geiger, H., 394, 400, 404
Donnell, L. H., 180, 482, 526 Geil, G. H., 180, 274
Dorn, J. E., 213, 257, 269, 352 Geiringer, H., 152, 404, 538
Dreyer, A., 285 Gensamer, M., 287
Duguet, C, 219, 314, 406, 467, 557 Gibbs, J. W., 151, 152, 162
DuPont, 300 Glasstone, S., 41, 199
Durand, W. F., 152, 400 Glathard, J. L., 188, 189
Glcyzal, A. N., 404, 419
E Glocker, R., 38
Goekowski, S., 297
Eddy, R. P., 510, 511 Goodier, N., 398, 477
Edgerton, H. E., 193 Gorens, P., 45
Edwards, S. H., 252 Gough, H. J., 53, 200
Eiehinger, A., 238, 251, 286 Graf, O., 96
Elam, C, 50-53, 55, 277, 297, 316 Green, 151
(See also Tipper, Mrs. C.) Greenfield, M., 192
Enger, 499 Greenspan, M., 321, 322
Espey, G., 261 Greenwood, J. N., 191
Euler, L., 373, 530 Griffis, LeVan, 262-264, 267
Evvald, P. P., 38 Griffith, A. A., 35, 40, 63, 196-200, 499
Eyring, H., 41, 68, 69, 199, 447 Griggs, D. T., 244
Grimison, E. D., 477
Guest, J. J., 220, 239, 248, 406, 409
Gutenberg, B., 423
Faraday, M., 275, 527
Farren, W. S., 51, 55 H
Fell, E. W., 275, 287
Fettweis, 285 Haigh, B. P., 208
Fiek, G., 70 Hanemann, H., 285
Fischer, F. P., 367 Hargreaves, M. E., 55
Fischer, J. C, 78, 441, 455, 467 Hartmann, L., 218, 275, 525
Flanigan, A. E., 83, 85, 185 Hartmann, W., 543
Flint, R. F., 223 Hausser, K. W., 39
Fliiggc, W., 394 Heidenreich, R. D., 53
Foppl, A., 177, 238, 239, 509 Heiser, G. H., 25
Foppl, L., 545 Helmholtz, H., 151
AUTHOR INDEX 5G3

Hencky, H., 107, 209, 230, 386, 482, Kelvin, Lord, 151
545 Keulegan, G. H., 29
Hensel, R., 285 Khristianovith, S., 546
Herbert, 347, 358 Kiesskalt, 32
Hetenyi, M., 179, 394 Kirchhoff, G., 151
Hettinger, L. H., 192 Kirkaldy, 299
Heyn, 275 Kirech, B., 289
Hill, R., 257, 319, 327, 352, 425, 437, Knopf, A., 225
Hodge, P. G., 404, 505 Korber, F., 21, 31, 43, 73, 285, 297, 319,
Hohenemser, K., 435 456, 539
Holmquist, J. H., 251 Koster, W., 284
Hooke, R., 76, 236, 353, 381, 422 Kotter, F., 543
Hort, H., 55, 400, 404 Kriiger, W., 538
Houwink, R., 29, 200, 404 Kusnetzov, 183
Hrennikoff, A., 359
Huber, M. I., 209, 230 L
Hiibers, K., 339, 340
Huckel, E., 40 Laidler, K., 41, 199
Hudson, G., 192 Lamb, H., 151, 400
Hutchings, P. F., 50 Lambert, J. L., 205, 343
Lame, 389, 452, 458
I Landgraf, F. K., 196
Lankford, W. T., 77, 261
llyushin, A. A., 85, 107, 386, 404, Laszlo, F., 482
420, 424 Latter, A. G., 213, 352
I.shlinsky, A. J., 85, 420 Leaderman, H., 29, 42
Hereon, F. K. T., 87, 404 Lee, A. R., 419, 425, 437, 461
Lee, G. H., 477
J Lehmann, 40
Leibenson, L. S., 505
Jeans, J. II., 30 Leiter, R. W., 297, 311, 312
Jeansen, C. F., 467 Leon, A., 182, 218, 221, 222, 408, 558
Jeffreys, H., 15, 423 Lessells, J. M., 238, 251
Jellinek, 213 Levistkaia, 178
Jenkins, W. H., 180 Levy, M., 386, 427
Jenks, G. F., 186 L'Hermite, R., 350, 352, 394, 419
Jensen, V. P., 359 Liepmann, H. W., 552
Jevons, I. D., 275, 287, 525 Lihl, F., 286
Joffe, A. F., 178, 195, 200 Lihotzky, 280
Johnson, A. E., 417 Lode, W., 20, 238, 239, 244-246, 248, 250,
Jones, H., 199 253
Lohan, R., 554
K Longwell, C. R., 223
Losco, 10. F., 47
Kachanov, L. N., 404 Losenhausen, 246, 247, 429
Karmdn, T. von, 61, 177, 238-240, Love, A. E. H., 119, 151, 166, 169, 394,
371, 374, 375, 552 491
Kasik, G. B., 261 Low, J. R., 289
Kauzmann, W., 69, 417 Lowenstein, L. C, 482
Kavanngh, J. C, 375 Lucretius, 19
5(34 THEORY OF FWiV AND FRACTURE OF SOLIDS

Lilders, 275, 276, 270-282, 288, 295, 320, X


328, 525
Ludwik, P., 21, 68, 72, 73, 81, 130, 238, Navier, M. H., 151, 388, 390, 399
250, 252, 285, 316, 343, 349, 409 Nehl, F., 399-341
Neuber, H., 179, 394
M
Newman, S. B., 321, 322
McAdam, I). J., 180-182,

I.,
222, 238, 274 Newton, 421
McCullough, G. H., 358 Niggli, 61
MacDougall, 1). P., 36 Ninninger, 281
MacGregor, C. W., 78, 81, 180, 238, 251, Norton, F. H., 25, 255, 257
271, 285, 298, 349, 437, 441, 455, 456, Nowick, A. S., 69
467, 489 Nye, J. F., 61
Machin, E. 8., 69
Macrea, A. C, 467

0
McVetty, P. G., 417
Maier, A., 238 Ohcrhoffer, P., 43, 515
Manjoine, M. J., 13, 47, 83, 85, 179, 194, Odquist, F. K. G., 417, 489
205, 298, 312-314, 343 Orowan, E., 62, 63, 184, 200
Mann, H. C, 186 Osgood, W. It., 375
Marburg, E., 504
Marcus, H., 394
Marin, J., 179
Mark, H., 42 Palm, J. H., 286
Martens, 275, 299 Parker, E. R, 83, 85, 185, 265, 267, 270
Maxwell, J. C, 151, 275, 421, 422 Pearson, K., 405, 527
Mebs, It. W., 180 Peck, V. G., 53
Mehl, R. F., 43, 53 Piccard, A., 504
Mehldahl, A., 208, 220 Piobert, 287
Meyer, Eugen, 21, 353 Pointing, J. H., 350
Meyer, H., 339-341 Poisson, S. D., 76, 151, 209, 380, 383,
Miklowitz, J., 82, 283, 298, 301, 303, 309, 387, 424, 458, 470
317, 327 Polanyi, M., 51, 63
Miller, C. P., 252 Pomp, A., 21, 343
Mises, R. von, 92, 104, 230, 249, 386, 403, Poncelet, E. F., 193, 198
545, 552, 554 Poulter, T. C, 35, 36, 177
Mohr, O., 3, 94, 96, 99, 106, 115, 151, 182, Prager, W., 404, 405, 410, 415, 435, 450
185, 214, 217, 224, 275, 325, 406, 407, Prandtl, L., 21, 27, 57, 58, 62, 144, 152,
557, 558 200, 204, 280, 330, 336, 339, 347,
Mollendorff, W., 82 393, 400, 407, 417, 429, 433, 491,
Moore, H. F., 53, 54, 67 493, 534, 536, 543, 545, 552, 555,
Morikawa, G. K., 262-264, 267 557
Morrison, J. L. M., 297 Preston, F. W., 188, 189
Morsch, E., 207, 375 Puckett, A. E., 552
Moser, M., 297 Pugh, E. M., 36
Mostov, W., 404 Putnam, W. I., 238
Mott, N. F., 199
Mugge, 61
0

Muller, A., 38, 39


Midler, W., 400 Quinney, H., 55, 238, 248, 2.50. 253. 297,
Munroe, 36 435
AUTHOR INDEX 565

R Seitz, F., 43, 51, 56, 199


Shaw, F. S., 507, 508, 510, 511
Rankine, 222, 556 Shevandin, E., 190
Raylcigh, 13, 151 Siebel, E., 21, 79, 238, 297, 309, 319, 339,
Read, T. A., 51 341, 343, 345, 539
Reiner, ML, 15, 404, 419, 421 Sinitski, A. K., 437
Reissner, II., 454, 543 Smekal, A., 62, 64, 200
Rejto, A., 71 Smith, J. B., 266
Rendulic, L., 221 Smith, S. L., 56
Reuss, E., 429, 434, 525, 526 Sokolnikoff, I. S., 394
Richart, F. E., 238 Sokolowsky, W. W., 85, 404, 437, 505,
Riedel, W., 339, 537, 538 556
Riemann, 546 Sommerfold, A., 552
Rigden, P. 0., 419 Sonntag, A., 246, 429
Rinne, F., 204, 330 Sopwith, C. R., 467
Robinson, E. L., 489 Southwell, R. V., 394
Roderick, J. W., 375 Spencer, C. D., 13
Ros, M., 238, 251

I.,
Spiridonova, N. 80, 85
Rose, B. A., 20 Stang, A. H., 321, 322
Rudeloff, M., 299 Stepanow, A. W., 195
Russel, R. S., 194 Sternberg, E., 522
Stodola, A., 482, 504
Stokes, 151, 399
Stone, M., 280
Sachs, G., 43, 44, 70, 261 Storp, H., 21
Sack, R. S., 21 Strauss, B., 279, 367
Sadowsky, M. A., 505, 522, 546 Struth, W., 193
Saibel, E., 77, 202, 261 Swainger, K. II., 419, 425
St. Venant, B., 151, 208, 386, 403, 405, Swift, II. W., 350-352
527-530
Schardin, II., 193
Seheel, C, 394, 400, 404
Scheu, R., 252, 285, 343, 349 Tammann, G., 31, 40, 43, 419
Schiebold, E., 515 Tangerdink, W., 285
Schilling, F., 498 Tapsell, II. F., 417
Schlechtweg, H., 393 Taylor, Sir Geoffrey, 36, 51, 53, 55, 59,
Schleicher, F., 227 62, 69, 200, 238, 248, 250, 253, 275,
Schmid, E., 44, 51, 63, 184 305, 435, 499
Schmidt, Erhard, 545 Taylor, Nelson W., 188
Schmidt, R., 253 Tertsch, 183
Schockley, W., 53 Thompson, A. S., 489
Scholl, 278 Thomsen, E. G., 213, 269
Scholz, P., 39 Tierney, W. D., 489
Schossow, G. J., 477 Tictjcns, O., 144, 400
Schrader, A., 285 Timoshenko, S., 288, 397, 491, 509
Schwaigercr, S., 297 Tipper, Mrs. C., 51, 184, 266
Scott, H., 47, 179 Todhunter, I., 405, 527
Secly, F. B., 238 Toliner, W., 504
Seidl, E., 515 Topler, 280
Seigle, M. J., 517 Torre, G, 216, 228, 407, 408, 558
566 THEORY OF FLOW AND FRACTURE OF SOLIDS

Trefftz, E., 504, 519 Wever, F., 31


Tresca, 211, 405, 527 Widmannstiidten, 281
Trevelyan, R. C, 10 Willis, B., 223
Troxell, G. E., 265, 267, 270 Willis, R., 223
Trumpler, W. E., 205 Wilson, E. B., 152, 162
Tupper, S. J., 425, 437, 461 Wilson, R. O., 36
Turner, T. H., 275, 287, 525 Windenburg, D. F., 187
Winlock, J., 297, 311, 312
V Winzer, A., 415, 450
Wittmann, F., 190
Uffelmann, L., 36 Wood, W. A., 56, 200
Wordcn, C. O., 196
Wyckoff, R., 38
Ver, T., 53, 54
Voigt, W., 162

W Yensen, T. D., 186


Young, A. P., 179
Wahl, A. M., 289, 291
Weibull, W., 191
Welch, L. E., 78, 456
Welter, G., 297, 309 Zappfe, C. A., 196
Westergaard, H. M., 208, 375 Zwicky, F., 63, 64, 199, 200
SUBJECT INDEX
A Cohesion of liquids, 35
Collapse of steel tubes, 251
Active plastic state, 535 Compatibility, condition of, 390, 403
Affine transformation, 157, 170 Compressibility, of glasses, 34
Affinor, 155 of liquids, 34
Aluminum, necking of, at 600°C, 83 of metals, 32
stress-strain curve of, 73 Compression tests, 328, 341
Amorphous solids, 12 of porcelain, 343
Anisotropic state of matter, 38 of steel, 343
Antecedent in dyad, 155, 157 Conchoidal fracture, 196
Antisymmetric dyadic, 159 Conjugate dyad, 155
Apparatus for torsion analogy, 500 Consequent in dyad, 155, 157
Arrangement, of atoms, in chains, 40 Constant load-rate test, 311
inordered, 38 Constant slope, surface of, 495
ordered, 38 Constant strain-rate test, 313
of inclusions in rocks, 223 Constrained flow, 426
of joints in rocks, 223 Continuity, condition of, 396
Atom rows, potential energy of, 60 Conventional strain, 17, 74
Atoms, lattice of, 39 Conventional stress-strain curve, 17, 70
ordered arrangement of, 38 Copper, brazing of, 47
Axes of strain, principal, 148 creep of, 25
flow tests of, 256
B stress-strain curve of, 18, 20, 72
tensile fracture of, 82
Bauschinger effect, 20 Cracks, fatigue, 54
Bituminous solids, 12 in glass, 192
Branch lines, 546 propagation velocity of, 192
Brazing of copper, 47 in rocks, 223
Breakdown of grain structure, 49, 65, 67 Creep, 4, 24
Brinell hardness, 19 of candles, 12
Brittle fracture, 178, 181, 187 of copper, 25
Brittle material, 70 of glass, 13
Brittle state, 207 steady, theory of, 416
Bulk modulus, 380 of steel, 25, 257
Criteria of failure, 175
C Crystalline fracture, 182
Crystalline structure of metals, 43
Characteristics, of partial differential Cubic cleavage, 183, 186
equations, 547, 552 Cubical dilatation, 116, 131
Circles, as envelopes of slip lines, 539 Cup and cone fracture, 81, 82, 183, 185
Cleavage cracks in rocks, 223 Curvilinear coordinates along slip lines,
Cleavage fracture, 183, 190, 202 543
Coefficient of viscosity, 11, 386, 396 Cycloids as slip lines, 535
508 THEORY OF FLOW AND FRACTURE OF SOLIDS

Cylinder, of ideally plastic material, 436 Envelope, of Mohr's circles, 215


plastic flow in, 436, 452 of slip lines, 546
rotating, 482 Equiareal curves, 540
of strain hardening metal, yielding of, Etching figures, 39
448 Etching method, Fry, 279
Etching pits, 185
D Expansion of hole in plate, 476
Extension, 70
Deformation, permanent, 17 Extensometer, 74, 75
work of, 78
Deviator of stress, 105 F
Dilatation of volume, 116, 131
Disks, overspeeding of, 482 Failure, criteria of, 175
plastic flow in, 472 Fatigue cracks, 54
rotating, 485 Faults, geological, 223
Dislocation, theory of, 50 Finite shearing strain, 121
Displacement, components of, 109, 115 Finite strain, 117, 163
field of, under uniform stress, 547 nonhomogeneous, 160
Displacement vector, 154 Flaws, effect of, on strength, 190
Drop of load, 298 Flow, constrained, 426
Ductility of metals, 3, 70, 250 radial, 538
Dyad, 155 Flow figures, 275, 281, 283, 330, 367
antecedent in, 155, 157 Flow function, for shear, 386
conjugate, 155 for tension, 386
consequent in, 155, 157 Flow layers, 276, 294, 328, 512
Dyadic (binomial), 155, 156 effect of nitrogen on, 284
antisymmetric, 159 Fluid state, 11
matrix of, 158, 163 Fluidal structure, 276, 294, 328, 512, 515,
nonion form of, 158 524
normal form of, 158 Fracture, brittle, 178, 181, 187
symmetric, 159 cleavage, 183, 190, 202
complex, 204
E conchoidal, 196
conditions of, 175, 178, 181
Earth pressure, theory of, 225 of copper, 82, 256
Elastic aftereffect, 28 crystalline, 182
Elastic deformation, 17 cup and cone (see Cup and cone frac
Elastic hysteresis, 26, 27, 58 ture)
Elastic solid, theory of, 388 effect of size on, 191
Elastic strain energy, 20!) of glass, 189, 192
Elastic strain theory, maximum, 208 along grain boundaries, 193
Elastic torsion, 491 Griffith theory of, 196
Elasticity, 14 herringbone (sec Herringbone fracture)
law of, 236 as instability of equilibrium, 199
Elastico-viscous solid, 421 Leon theory of, 221
Electron microscope, observations of, 53 iMohr theory of, 214
Ellipse of plasticity, 213, 214, 473 of plastics, 195
Elliptic partial differential equations, 554 of porcelain, 187, 206
Elongation, quadratic, 117 shear (sec Shear fracture)
Elongation-shear ellipse, 125 of spherical tanks, 266
SUBJECT INDEX 569

Fracture, of steel under combined stress, Instability of equilibrium, fracture as, 199
213, 245, 259, 263, 265, 270 of stresses, in cylinders, yielding, 455
tensile (.see Tensile fracture) in plastico-viscous substance, 420
transcrystalline, 194 in tension bar, 305, 316
Fragmentation of lamellae, 200 of uniform yielding in twisted round
Fry etching method, 279 bar, 525
Invariants of stress tensor, 92
Isobars, in plane fields, 545
Isochromatics, 291, 292
Gauss' theorem, 169 Isoclines, in plane fields, 545
Geologic examples of flow of rocks, 5,
223
Glass, 12, 13, 41, 192
compressibility of, 34 Joint in rocks, 223
cracks in, 192
creep of, 13
Gradient in scalar field, 168
Grain boundary substance, effect of, on Lamellae, fragmentation of, 200
strength, 45, 285 Lamellar perlite, 46
Grain structure, breakdown of, 49, 65, 67 Lattice, of atoms, 39
Grooves, effect of, 517 Length change in twisted bars, 350
Limiting states of stress, 175
II Limiting surface, of rupture, 178
of yielding, 176, 220
Heat evolved in plastic straining, 53 Linear transformation, 111, 157
Helium, liquid states of, 40 Linear vector function, 114, 152
Herringbone fracture, 190, 191, 261, 266 Liquid state, 10
High hydrostatic pressure, behavior of Liquids, cohesion of, 35
matter under, 30 compressibility of, 34
effects of, on density, 30 Logarithmic spirals as slip lines, 539
on ice, 31 Logarithmic strain-rate law, 21, 22
on rigidity, 34 Luders' lines, 275
on water, 31 (See also Flow figures; Strain figures)
High polymers, 8
High-energy tests, 262 M
High-speed tensile tests, 23, 24
Hole in plate, expansion of, 476 Marble, flow tests for, 240
Homogeneous strain, 109 slip lines in, 330, 339
Hooke's law, 236 Matrix of dyadic, 158, 163
Hyperbolic partial differential equation, Mean normal stress, 102, 105
554 Mechanical similarity, 87
Hyperbolic sine speed law, 417 Membrane analogy, 491
Metal crystal, single, 44
Metals, compressibility of, 32
crystalline structure of, 43
Ice, varieties of, 31 ductility of, 3, 70, 250
Ideally plastic substance, 15, 401 Meteoric iron, 281
Idem factor, 156 Mohr envelope, theory of, 221
Imperfections, effect of, on strength, 196 Mosaic crystal, 65
Incompressibility, condition of, 74, 170 Munroe effect, 36

r
570 THEORY OF FLOW AND FRACTURE OF SOLIDS

X Plastic bending, 353


Plastic buckling, 371
Nabla operator, 167 Plastic collapse of tubes, 251
Natural strain, 74, 130 Plastic flow, in cylinders, 436, 452
Natural stress-strain curve, 73, 76 in disks, 472
Natural unit shear, 130, 139, 142, 143 rules of, 231
Necking, at constant rates of strain, 84 in turbine disks, 482
in copper bar, 82 Plastic mass, vortical flow in, 540
in flat bars, 326 Plastic region, spreading of, around holes,
at high temperature, 83 520, 522
in nylon filaments, 7, 300 Plastic state, active, 535
in stainless-steel bar, 84 of matter, 14, 207
in steel bar, 82 passive, 535
theory of, 80, 86, 271 Plastic stress function of torsion, 494
Nitrogen, effect of, on flow layers, 284 Plastic torsion, 347, 494
Notch effect, 194 Plasticity, condition of, 210, 403
Nylon, 6, 300 ellipse of, 213, 214, 473
of rock salt, 4
0 of rocks, 4, 5, 240
Plastico-viscous substance, 15, 418
Oblique fracture in tension, 319
Plastics, 6
Octahedral normal stress, 103
Polymers, 8
Octahedral planes, 38, 103
Polymorphism, 31
Octahedral shearing stress, 104, 210
Porcelain, compression tests of, 343
maximum, theory of, 210
fracture tests of, 187, 345
Octahedron, 103, 105
Pressure, high hydrostatic (see High
Orthogonal system of slip lines, 554
hydrostatic pressure)
properties of, 545
Principal axes of strain, 148
Overspeeding of disks, 482
Principal elongations, 119
Principal shearing stresses, 99
P
Principal strains, 148, 170
Paraffin, compression tests with, 286, 287, Principal stresses, 92
289, 331, 335, 337, 338 Propagation velocity of cracks, 192
Partial differential equations, character
R
istics of, 547, 552
Partial yielding, in flat rings, 472 Radial distribution of stress, 541
in heavy-walled cylinders, 436 Radial flow, 538
in revolving disks, 482 Rate of strain, 21, 309
in torsion, 497 Rate-process theory, 68
of hollow specimens, 507 Reciprocal strain diagram, 129
in twisted shaft of variable cross sec Reciprocal strain ellipsoid, 119
tions, 508 Recovery, 8, 26
Perlite, spheroidization of, 46 Recrystullization, 44
Photoelastic tests, 291 Refractalloy, 47
Plane plastic strain, 427, 543 Relaxation, 4, 15
Plane plastic stress, 556 Resultant strain, 383
Plane strain, 133, 390 Resultant stress, 161
Plane stress, 94, 390 Rhombic dodecahedron, 101
Planes, octahedral, 38, 103 Rhombus, unstretched, 110, 137, 145
of slip, orientation of, 217, 279 Riemann surface, 546
SUBJECT INDEX 571

Rocks, cleavage cracks in, 223 Solid, model of, by Bragg, 61

flow of, geologic examples of, 5, 223 by Prandtl, 57


inclusions in, arrangement of, 223 by Taylor, 59
joints in, arrangement of, 223 vitreous, 39
plasticity of, 4, 5, 240 Solid state, 10
Rotating cylinders, 482 Solids, amorphous, 12
Rotating disks, 485 bituminous, 12
Rotation, pure, 110 surface tension in, 197
Rupture, surface of, 178 Space derivative, 168
Space lattice, 39
S
Spheroidization of perlite, 46
Salt dome, 5 Spiral cracks in glass, 192
Sand hill analogy, 494 Spreading of plastic region around hole,
Sandstone, compression test with, 329 520, 522
Scalar, 152 Steel, compression tests of, 343
Scalar product, 153 constant load-rate tests for, 31 1
Schlieren method, 280 constant strain-rate tests for, 314
Shear, finite, 120, 123 creep of, 25, 257
flow function for, 386 flow tests of, under combined stress,
natural, 130, 139. 142 245, 259, 263, 265, 270
pure, 110, 133 fracture tests on, 190, 191
work under, 139
Steel tubes, collapse of, 251
simple, 109, 146
Strain, components of, 114
Shear fault in rocks, 223
conventional, 17, 74
Shear fracture, 182, 190, 202
finite (see Finite strain)
Shearing strain, 110, 117
homogeneous, 109
Shearing stress, 89
infinitesimal, 1 13
maximum, theory of, 211, 213
nonhomogeneous, 115
octahedral, 104, 210
natural, 74, 130
principal, 99
rate of, 21, 309
Silicone, 8
resultant, 383
Size, effect of, on fracture, 191
Strain ellipsoid, 119
Slenderness ratio, 364
Strain figures, 275
Slip, condition of, 217, 224
Slip bands, in copper aluminum, 52 in bending, 267, 283
in copper crystallites, 50 in compression, 276, 294, 328, 330
in crystals, 38, 49 in extruded tube, 346
in mild steel, 277 around hole, 281, 286, 289, 292
in pure iron, 53 in paraffin, 331
Slip lines, circles as envelopes of, 539 in plate, 282
curvilinear coordinates along, 543 in prisms, 336
cycloids as, 535 in rings, 334
envelope of, 546 in tension, 281, 282, 288, 322
logarithmic spirals as, 539 Strain hardening, of metals, theory of,
in marble, 330, 339 409

orthogonal system of, 554 tests on, 252


properties of, 545 theory of, 59
Soap film analogy, 491 Strains, principal, 148, 170
Solid, elastic, theory of, 388 Strength, effect of flaws on, 196
elastico-viscous, 421 ultimate, 70
572 THEORY OF FLOW AND FRACTURE OF SOLIDS
Strength theory, 207 T
of Coulomb, 219 Tensile fracture, 79, 82, 83, 183, 185
of Guest, 220 Tensile test, 70
of Huber-Hencky, 200 high-speed, 23, 24
of Mohr, 214 Tension, flow function for, 386
of St. Venant, 208 oblique fracture in, 319
Stress, deviator of, 105 triaxial, 178
limiting states of, 175 Tension-torsion machine, 247
maximum, theory of, 208 Tension-torsion tests, 248
mean normal, 102, 105 Tensor, 161
octahedral normal, 103 components of, 162
principal, 92 Torsion, 374
radial distribution of, 541 plastic, 347, 494
resultant, 383 plastic stress function of, 494
true, 71 Transcrystalline fracture, 194
Stress circle of Mohr, 94, 97 Triaxial tension, 178
Stress components, 89 True stress, 71

Stress concentration, 194, 415 Tubes, plastic collapse of, 251


Stress distribution, in cylinders, 460, 466 Twins in crystals, 49, 61
in disks, 476, 479 U
Stress ellipsoid, 92
Ultimate strength, 70
Stress function, 390, 399, 492, 494, 529,
Unit shear (see Shear)
531
Stress invariants, 92
Stress tensor, 161 Vector, 152
invariants of, 92 Vector product, 154
Stress-strain curve, 70 Vibrating string, 548
of aluminum, 73 Viscosity, 11, 12
conventional, 17, 70 coefficient of, 11, 386,396
of copper, 18, 20, 72 Viscous substance, theory of, 395
of mild steel, 18 Vitreous solid, 39
natural, 73, 76 Vortical flow in plastic mass, 540
of nylon, 303
w
peak of, 298
of wrought iron, 72 White dwarf star, 30
Work, of deformation, 78
Structure, of cast zinc, 517
under pure shear, 139
due to cooling, 515
Working length, 301, 318
of copper, 50
dendritic, 47
fluidal (see Fluidal structure) Yield point, lower, 298
insensitive properties, 64 of steel, 297
of rubber, 41 upper, 298, 307
sensitive properties, 64 Yield point elongation, 299, 312
of steel, 46 Yield stress, 13
Surface, of constant slope, 495 Yielding, partial (see Partial yielding)
of rupture, 178 surface of, 176, 220,221
of yielding, 176, 220, 221
Surface tension in solids, 197
Symmetric dyadic, 159 Zinc, cast, structure of, 517
14 USE DAY
RETURN TO DESK FROM WHICH BORROWED
ENVIRONMENTAL DESIGN
LIBRARY
This book is due on the last date stamped below, or
on the date to which renewed.
Renewed books are subject to immediate recall.

DEC p a 1070

DUE END RECALLABLE ONF WE TV ft—

PA
PnLL
II M n\/
NUV
n
9 ]; fb

'""'"MAR 2 6 1977

.iflN 2 5

REC'D OR FOUND

DFP * *— IwOl
O —1
L^fc.

LD 21A-50m-9,'67 General Library


University of California
(H5067sl0)476 Berkeley
UC BERKELEY LIBRARIES

CQ372b2TE0

r
I ■ ■■

Vous aimerez peut-être aussi