2-Plasticiy Theory and Application - A. MENDELSON - 3
2-Plasticiy Theory and Application - A. MENDELSON - 3
2-Plasticiy Theory and Application - A. MENDELSON - 3
APPLIED MECHANICS
F R E D LANDIS, EDITOR
PLASTICITY:
Theory and
Application
With the advent of the jet age followed closely by the space age, the theory
of plasticity has been brought forcibly into the forefront of engineering appli-
cation and design. Modern aircraft, missiles, and space vehicles must be
designed on the basis of minimum weight, which invariably means designing
into the plastic range to obtain maximum load to weight ratios. Moreover,
the facts of economic life have made the saving of material and more efficient
design a necessity for even the more earthbound industrial applications.
This book is the result of the author's teaching for several years of a
graduate course in plasticity for engineers at Case Institute of Technology.
It was soon realized that although a number of excellent books on plasticity
were available, none of them adequately met the requirements of the course.
The available books were either too theoretical and mathematical for the
average engineer and designer, or their main emphasis was placed on problems
of large plastic deformations such as are encountered in metal-forming
processes. Very little has been published in textbook form on the most
,important class of elastoplastic problems, in which the plastic strains are
of the same order of magnitude as the elastic strains, which are of such prime
'concc~rn to today's engineer. Furthermore, where such problems are treated,
usual assumptions of perfect plasticity are used, no attempt being made
take into account the strain-hardening properties of real materials.
A set of mimeographed notes was prepared which included the basic theory
placed primary emphasis on the solution of elastoplastic problems for
vii
viii Preface
Chapter 1. Introduction 1
xi
xii Contents
327
Chapter 14. Creep
327
14-1 BASIC CONCEPTS
331
14-2 MuLTIDIMENSIONAL PROBLEMS
333
14-3 UNIAXIAL CREEP IN INFINITE STRIP
335
14-4 CREEP IN ROTATING DISKS
347
Index
CHAPTER 1
INTRODUCTION
The history of plasticity as a science began in 1864 when Tresca [1] pub-
lished his results on punching and extrusion experiments and formulated his
famous yield criterion. A few years later, using Tresca's results, Saint-Venant
[2] and Levy [3] laid some of the foundations of the modern theory of plasti-
city. For the next 75 years progress was slow and spotty, although important
contributions were made by von Mises [4], Hencky [5], Prandtl [6], and others.
It is only since approximately 1945 that a unified theory began to emerge.
Since that time, concentrated efforts by many researchers have produced a
voluminous literature which is growing at a rapid rate. Brief but excellent
historical sketches are furnished by Hill [7] and Westergaard [8].
The theories of plasticity fall into two categories: physical theories and
mathematical theories. The physical theories seek to explain why metals flow
plastically. Looking at materials from a microscopic viewpoint, an attempt
made to determine what happens to the atoms, crystals, and grains of a
when plastic flow occurs. The mathematical theories, on the other
are phenomenological in nature and attempt to formalize and put into
form the results of macroscopic experiments, without probing very
into their physical basis. The eventual hope, of course, is for a merger
two approaches into one unified theory of plasticity which will both
the material behavior and provide the engineer and scientist with the
tools for practical application. The present treatise is concerned
the second of these categories, i.e., the mathematical theories of plasticity
1
2 Introduction [Ch. 1
and their application, as distinct from the physical theories. The latter belong
to the realm of the metal physicist or solid-state physicist.
We start by defining roughly and intuitively what is meant by a metal
flowing plastically. If one takes a thin strip of a metal such as aluminum and
clamps one end and applies a bending force to the other end, the end of the
strip will deflect. Upon removal of this force, if this force is not too large, the
end of the strip will spring back to its original position, and there will be no
apparent permanent deformation. If a sufficiently large load is applied to the
end, the end will not spring back all the way upon the removal of the load
but will remain permanently deformed, and we say that plastic flow has
occurred. Our objective in this case will not be to determine why the perma-
nent deformation took place but to describe what has happened in terms of
stresses, strains, and loads. Solutions of this particular problem can be found,
for example, in references [9] and [10].
In short, plasticity is the behavior of solid bodies in which they deform
permanently under the action of external loads, whereas elasticity is the
behavior of solid bodies in which they return to their original shape when the
external forces are removed. Actually, however, the elastic body is an ideali-
zation, because all bodies exhibit more or less plastic behavior even at the
smallest loads. For the so-called elastic body, however, this permanent
deformation is so small as to be practically not measurable, if the loads are
sufficiently small. Plasticity theory thus concerns itself with situations in
which the loads are sufficiently large so that measurable amounts of perma-
nent deformation occur. It should.further be noted that plastic deformation
is independent of the time under load. Time-dependent deformations are
discussed briefly in Section 2.4 and in Chapter 14.
The theory of plasticity can conveniently be divided into two ranges. At
one end are metal-forming processes such as forging, extrusion, drawing,
rolling, etc., which involve very large plastic strains and deformations. For
these types of problems the elastic strains can usually be neglected and the
material can be assumed to be perfectly plastic. At the other end of the scale
are a host of problems involving small plastic strains on the order of the
elastic strains. These types of problems are of prime importance to the
structural and machine designer. With the great premium currently placed
on the saving of weight in aircraft, missile, and space applications, the de-
signer can no longer use large factors of safety and "beef up" his design. He
must design for maximum load to weight ratio, and this inevitably means
designing into the plastic range. Even in more prosaic industrial applications
the competitive market is forcing the application of more efficient design.
In this book emphasis will be placed primarily on the second type of prob-
lem, i.e., the elastoplastic problems, where the plastic strains are of the same
Ch. 1] Introduction 3
References
1. H. Tresca, Sur l'ecoulement des corps solids soumis a de fortes pression,
Compt. Rend., 59, 1864, p. 754.
2. B. de Saint-Venant, Memoire sur l'etablissement des equations differen-
tielles des mouvements interieurs operes dans les corps solides ductiles au
dela des limites oil l'elasticite pourrait les ramener a leur premier etat,
Compt. Rend., 70, 1870, pp. 473-480.
3. M. Levy, Memoire sur les equations generales des mouvements interieurs
des corps solides ductiles au dela des limites oil l'elasticite pourrait les
ramener a leur premier etat, Compt. Rend., 70, 1870, pp. 1323-1325.
4. R. von Mises, Mechanik der festen Koerper im plastisch deformablen
Zustant, Goettinger Nachr., Math.-Phys. Kl., 1913, pp. 582-592.
5. H. Hencky, Zur Theorie plastischer Deformationen und der hierdurch im
Material hervorgerufenen Nebenspannungen, Proceedings of the 1st Inter-
national Congress on Applied Mechanics, Delft, Technische Boekhandel en
Druckerij, J. Waltman, Jr., 1925, pp. 312-317.
6. L. Prandtl, Spannungsverteilung in plastischen Koerpern, Proceedings of the
1st International Congress on Applied Mechanics, Delft, 1924, pp. 43-54.
7. R. Hill, The Mathematical Theory of Plasticity, Oxford University Press,
London, 1950.
8. H. M. Westergaard, Theory of Elasticity and Plasticity, Harvard University
Press, Cambridge, 1952.
B. W. Shaffer and R. N. House, The Elastic-Plastic Stress Distribution
Within a Wide Curved Bar Subjected to Pure Bending. J. Appl. Mech., 22,
No. 3, 1955, pp. 305-310.
B. W. Shaffer and R. N. House, Displacements in a Wide Curved Bar,
J. Appl. Mech., 24, No. 3, 1957, pp. 447-452.
CHAPTER 2
BASIC
EXPERIMENTS
The simplest and most common experiment, as well as the most important,
is the standard tensile test. A cylindrical test specimen such as shown in
Figure 2.1.1 is inserted into the tensile machine, the load is increased, and the
readings of the load, the extension of the gage length inscribed on the speci-
men, and/or the decrease in diameter are recorded. A typical load extension
diagram is shown in Figure 2.1.2.
The nominal stress, defined as the load divided by the original cross-
sectional area, is plotted against the conventional or engineering strain, defined
as the increase in length per unit original length. Nominal stress is represented
by
p
a=-
n Ao
4
Sec. 2-1] Tensile Test 5
p
FIGURE 2.1.1 Tensile specimen.
Initially the relation between stress and strain is essentially linear. This linear
part of the curve extends up to the point A, which is called the proportional
limit. It is in this range that the linear theory of elasticity, using Hook's law,
is valid. Upon further increase of the load, the strain no longer increases
linearly with stress, but the material still remains elastic; i.e., upon removal
of the load the specimen returns to its original length. This condition will
prevail until some point B, called the elastic limit, or yield point, is reached.
In most materials there is very little difference between the proportional limit
A and the elastic limit B. For our purpose, we shall consider them to be the
same. Furthermore, the values of these points depend on the sensitivity of the
measuring instruments. For some materials the yield point is so poorly defined
that it is arbitrarily taken to be at some fixed value of permanent strain, such
as 0.2 per cent. The stress at this point is usually called the offset yield strength,
or the proof strength. Beyond the elastic limit, permanent deformation, called
plastic deformation, takes place. The strain at the elastic limit is of the order
of magnitude of 0.001, or 0.1 per cent. As the load is increased beyond the
elastic limit, the strain increases at a greater rate. However, the specimen will
not deform further unless the load is increased. This condition is called work
hardening, or strain hardening. The stress required for further plastic flow is
called flow stress. Finally a point is reached, C, where the load is a maximum.
Beyond this point, called the point of maximum load, or point of instability, the
specimen" necks down" rapidly and fractures at D. Beyond C a complicated
triaxial state of stress exists. The point C therefore represents the limit of the
useful part of the tensile test as far as plasticity theory is concerned. The stress
at the maximum load point Cis called the tensile strength, or ultimate stress.
If at any point between the elastic limit B and the maximum load point C
the load is removed, unloading will take place along a line parallel to the
elastic line, as shown in the figure by B' C'. Part of the strain is thus recovered
and part remains permanently. The total strain can therefore be considered
as being made up of two parts, ee, the elastic component, and eP, the plastic
component:
(2.1.3)
Upon reloading, the unloading line, B'C', is retraced with very minor
deviations. Actually a very thin hysteresis loop is formed, which is usually
neglected. Plastic flow does not start again until the point B' is reached. With
further loading, the stress-strain curve is continued along B'C as if no
unloading had occurred. Point B' can thus be considered as a new yield point
for the strain-hardened material.
A few materials, such as annealed mild steel, exhibit a sharp drop in yield
after the upper yield point B is reached, as shown by the dashed line. The
specimen will then extend at approximately constant load to a strain of about
10 times the initial yield before the load will start increasing again as the
material begins to work harden. The flat portion of the curve, called the
lower yield, actually represents an average of a series of unstable jumps
between the upper and lower yields caused by the propagation of Luder bands
across the specimen. The upper yield point is very sensitive to small bendin~
Sec. 2-2] True Stress-Strain Curve 7
We have discussed the plot of the nominal stress versus the conventional
strain. It is evident, however, that this nominal stress is not the true stress
acting in the specimen, since the cross-sectional area of the specimen is
decreasing with load. At stresses up to and near the yield, this distinction is
of no importance. At higher stresses and strains this difference becomes
important. The true stress can readily be obtained from the nominal stress
as follows. If small changes in volume are neglected, i.e., the material is
assumed to be incompressible, then
Aolo =AI
where A 0 and /0 are the original cross-sectional area and gage length and
A and l are the current values. If P is the load, then the true stress a is
P PI
a=-=-
A A 0 l0
dl
de= T (2.2.2)
the total strain in going from some initial length /0 to the length l is
(2.2.3)
8 Basic Experiments [Ch. 2
e = ln (1 + e) (2.2.4)
For small strains the two are practically identical, and for most problems
considered the conventional strain will be used. The natural strain, however,
has several advantages. Natural strains are additive, but conventional strains
are not. Second, if a ductile material is tested in compression and in tension,
the true-stress versus true-strain curves are almost identical, whereas they are
quite different if conventional strain is used. Finally, the incompressibility
condition to be used later becomes simply
(2.2.5)
P = aA
dP = a dA + A da = 0
da
or
a
Also A 0 l0 =AI
Ad!+ IdA= 0
Sec. 2-2] True Stress-Strain Curve 9
dA dl
or A -~
da dl
Hence -=-=de
a l
da
or -=a
de
da a
(2.2.7)
de = 1 +e
On a plot of a versus e, the value of a at which the load is a maximum
occurs where the slope is equal to the stress; i.e., one must draw a tangent to
that point of the curve for which the subtangent is equal to 1, as shown in
Figure 2.2.1. Discussions of the stress-strain curve and the strain distributions
in the neck of a tensile specimen after necking has started can be found in
"'""·l'-'11~·"~ [2] and [3].
Alternatively, the true stress-strain curve can be obtained by measuring the
ctttamt>tr.al strain rather than the longitudinal strain, provided the tensile sped-
has a circular section. Thus, if en is the strain in the diametral direction,
D- D 0
en= Do (2.2.8)
D 0 is the initial diameter and Dis the diameter at the true stress, a. The
diametral strain is
(2.2.9)
10 Basic Experiments [Ch. 2
e- Do
= - 2-Bv = 21nD (2.2.10)
The true strain at any load can therefore be determined by measuring the
change in diameter of the specimen.
From equation (2.2.10) it is seen that the true strain can also be written
A
e = ln_Q (2.2.11)
A
The quantity on the right of equation (2.2.11) is called the true reduction in
area. Equation (2.2.11) states that the true strain is equal to the true reduction
in area.
Figure 2.2.2 (from reference [4]) shows the true stress-strain curves for a
320
"'r0
)(
0.80 1.00
Natural Strain- E
1,000r------------------------,
~
+-
(/)
Q)
::l
~
o Necking
x Fracture
1.0
Natural strain 8
variety of materials. The ends of the curves represent the points of fracture
and the circle on each curve represents the maximum load point or instability
point for that curve. To show the complete curves to fracture, the abscissa is
such that the elastic parts of the curves are too small to be seen. These curves
are also shown replotted on log-log coordinates in Figure 2.2.3. Note that
most of the curves appear as straight lines on this log-log plot. This indicates
that they can be represented by an equation of the form
a= Aen (2.2.12)
where A and n are material constants with n the slope of the curve when
plotted on log-log coordinates. A is called the strength coefficient and n is
called the strain-hardening exponent.
It follows from equation (2.2. 7) that for a material which behaves according
equation (2.2.12), the true strain at the point of maximum load is given by
e=n (2.2.13)
simple relation (2.2.13) has been found useful in fracture studies. It also
a simple method for determining the instability point on the true
curve .
.12q1uatton (2.2.12) will, of course, not fit all materials, nor will it be valid
all strains or very large strains. However, Marin [5] has studied 31
12 Basic Experiments [Ch. 2
different materials and found that the average deviation between the theoreti-
cal values of e as given by equation (2.2.13) and the actual values was 2 per
cent.
A single quantity which represents the ability of a material to deform
plastically is the ductility of the material. The most common measure of
ductility is the per cent elongation in the tensile test, i.e., the per cent strain at
fracture. Thus, if 11 is the gage length at fracture and /0 is the initial gage
length, then the per cent elongation is
I,- lo
e1 = - - X 100 (2.2.14)
lo
Together with the per cent elongation as given by equation (2.2.14), one
must also specify the initial gage length /0 , since the per cent elongation will
depend on the gage length used. This is due to the fact that once the instability
point is reached and necking starts, most of the deformation occurs in the
smallest cross section, with only a relatively small amount of additional de-
formation occurring throughout the rest of the gage length. The longer the
gage length used, the smaller the per cent elongation will be. The ductility is
therefore reported as the per cent elongation for a given gage length.
A better measure for ductility, however, is the true strain at fracture:
(2.2.15)
where A 0 is the initial area and A 1 is the area at fracture. As mentioned pre-
viously, ductility is an important property which measures the ability of a
material to deform. A material with low ductility cannot deform very much
under load and will behave in a brittle fashion. Unexpectedly high loads may
cause such a material to fracture, whereas a material with high ductility will
deform under similar loads without fracturing. A cyclic load above the
will cause a low-ductility material to fail in relatively few cycles, whereas
high-ductility material will fail after a much larger number of cycles (at
for very low cycle fatigue). In metal-forming processes such as rolling,
ing, forging, etc., a sufficient amount of ductility is needed to prevent
during the forming process.
Sec. 2-3] Compression Test and the Bauschinger Effect 13
If instead of a tensile test one runs a compression test and plots nominal
stress against conventional strain, a different curve will be obtained than for
the tensile test. However, if the true stress is plotted against the true strain,
practically identical curves are usually obtained. The yield points in tension
and compression will, for example, generally be the same. If, however, a
metal is first deformed by uniform tension and the load is removed and the
specimen is reloaded in compression, the yield point obtained in compression
will be considerably less than the initial yield in tension. This has been ex-
plained as being the result of the residual stresses left in the material due to
the tensile deformations [6]. A perhaps better explanation is based on the
anisotropy of the dislocation field produced by loading [7]. This effect is
called the Bauschinger effect, and is present whenever there is a reversal of the
stress field. The Bauschinger effect is very important in cyclic plasticity
studies. Unfortunately, however, it enormously complicates the problem and
is therefore usually ignored.
There are several simplified models used to describe the Bauschinger effect.
These are illustrated in Figure 2.3.1 (from reference [8]). At one extreme it is
assumed that the elastic unloading range will be double the initial yield stress.
If the initial yield stress in tension is a 0 , then the specimen will yield in com-
pression after being stressed in tension to a = a 1 when
This is shown as path ABCDE in Figure 2.3.1. According to this theory, then,
the total elastic range of the material remains constant, the initial compressive
yield being reduced by the same amount as the tensile yield is raised.
At the other extreme there is isotropic hardening. This theory assumes
the mechanism that produces hardening acts equally in tension and
cornp1res1dOJ11. Thus compressive yielding will occur when
shown by the path ABCFG. This is the simplest of the theories to apply and
the one most frequently used.
'£e:tw1~P.n these theories there is a theory which assumes that the tensile and
mJ>re1>Sl,,e yields are independent of each other. The compressive yield
is independent of the amount of tensile hardening and remains at
a= - a0
14 Basic Experiments [Ch. 2
G
FIGURE 2.3.1 Theories for Bauschinger effect.
Tests on the effect of the rate of straining and of temperature on the prop-
erties of mild steel were carried out by Manjoine [9], among others. The effect
of increasing the strain rate is generally to increase the tensile yield, as shown
in Figure 2.4.1. For materials with a lower yield, such as mild steel, the stress-
~---10 3 sec-1
~----10- 2 sec- 1
e
FIGURE 2.4.1 Effect of strain rate.
strain curve may approach that of a perfectly plastic material. For other
materials the reverse will be true, and the strain hardening will increase with
strain rate [16]. These effects are important in some metal-forming processes
which are performed at very high strain rates. These types of processes will
not be discussed in this text.
Temperature has a very important effect on metal properties. At very low
temperatures metals which are very ductile can become very brittle. This is
illustrated in Figure 2.4.2 (from reference [10]). The temperature at which the
ductility changes so rapidly is called the transition temperature. Such strong
temperature and strain-rate effects occur more generally in metals with
body-centered-cubic structures.
100
0~
0Q) 80
0 100min-1
.s 60
c
0
:;::: 40
u
::J
"0
Q)
20
0:
At the other end of the time and temperature scales is the phenomenon of
creep. Creep is a continuous deformation with time under constant load and
occurs primarily at high temperatures, although some metals, e.g., lead, will
creep at room temperature. Although it is questionable whether plasticity
theories can be applied to the creep phenomenon, it is the usual practice to
do so and, in Chapter 14, we shall describe how this is done. A typical set of
creep curves is shown in Figure 2.4.3.
Time
FIGURE 2.4.3 Creep curves.
r--p
(a)
t
cr
0
7tan- 1 mE p
_i__.___ _ _ __
(c)
(d)
(e)
u =OP
1-
0 e =Iii OR
(a)
T----'---------,--------'------T
(b) IR Pe
T-----'-------~-------L------T
(c) IR
T------~----------.--------~-----T
(d) IR PI!
T-------''--------,---------'----T
(e) IR p.
T----------~---,------------L-T
(f)
(g)
(h) IR
T----------~----------~-----------------'-------T
(i) IR
T-------~----------~-----------------'-------T
0
FIGURE 2.6.2 Kinematic model.
Sec:. 2-6} Idealization of the Stress-Strain Curve 19
A~----+------------+F~-----
(2.7.1)
(2.7.2)
(2.7.5)
where e is the base of natural logarithms, e0 the yield strain, a 0 the yield stress,
E the elastic modulus, and m, n, k, a, b, and c constants.
It is also possible to fit the plastic part of the stress-strain curve as
ately as desired by a polynomial of arbitrary degree, i.e.,
where e0 is the yield strain. For linear strain hardening all the CO(~ffilcieJilf
beginning with a2 are zero.
Ch. 2] Problems 21
Problems
1. Show that natural strains are additive whereas conventional strains are not.
2. Assume that a material behaves elastically up to the point of instability.
Show that the natural strain at this point is unity.
3. Derive equation (2.2.13).
4. Let the stress-strain curve of a material be given by a = Aen, where e is the
conventional strain. Show that at the point of instability
n
e=--
1- n
5. In a standard tensile test using a t-in.-diameter specimen with a l-in. gage
length, the following data were obtained. At a load of 10,000 lb, the conven-
tional strain was 0.10, and at a load of 12,000 lb, the conventional strain was
0.60. Find the true stresses and strains for these two conditions. Determine
the strength coefficient A, the strain-hardening exponent n, the change in
gage length at the maximum load, and the maximum load assuming equation
(2.2.12) to hold.
6. A tensile load is applied to a thin-walled hollow circular cylinder. Determine
the change in wall thickness and in mean radius at the point of maximum
load, if the stress-strain curve is given by a = Aen, where e is the conven-
tional strain and a is the true stress.
7. Derive the incompressibility conditions (2.2.5) and (2.2.6).
8. The following data were obtained in a tensile test on a 0.505-in.-diameter
specimen:
by W, the friction coefficient by p,, and the force by P. For example, for the
first model, the equation of the stress-strain curve is a = Ee and the corres-
ponding model equation is P = kx. Thus
11. For the kinematic model of Figure 2.6.2, show that e = ORfm.
12. Describe a kinematic model similar to that shown in Figure 2.6.2 for iso-
tropic hardening.
13. Sketch typical stress-strain curves that would be obtained using Ludwik's
expression for the following cases:
(a) n = 1.
(b) 0 s n < 1.
(c) ao = 0, n = 0, t, 1.
References
1. P. Ludwik, Elemente der technologischen Mechanik, Springer, Berlin, 1909.
2. J. D. Lubahn and R. P. Felgar, Plasticity and Creep of Metals, Wiley, New
York, 1961.
3. G. E. Dieter, Jr., Mechanical Metallurgy, McGraw-Hill, New York, 1961.
4. H. Schwartzbart and W. F. Brown, Jr., Notch-Bar Tensile Properties of
Various Materials and their Relation to the Unnotch Flow Curve and
Notch Sharpness, Trans. ASM, 46, 998, 1954.
5. J. Marin, Mechanical Behavior of Engineering Materials, Prentice-Hall,
Englewood Cliffs, N.J., 1962.
6. R. Hill, The Mathematical Theory of Plasticity, Oxford Univ. Press, London,
1950.
7. D. Mclean, Mechanical Properties of Metals, Wiley, New York, 1962.
8. J. N. Goodier and P. G. Hodge, Jr., Elasticity and Plasticity, Wiley,
York, 1958.
9. M. J. Manjoine, Influence of Rate of Strain and Temperature on
Stresses of Mild Steel, J. Appl. Mech., 11, A-211, 1944.
10. A. W. Magnusson and W. M. Baldwin, Low Temperature Brittleness,
Mech. Phys. Solids, 5, 172, 1957.
11. P. W. Bridgman, The Effect of Hydrostatic Pressure on the Fracture
Brittle Substances, J. Appl. Phys., 18, 246, 1947.
12. P. W. Bridgman, Studies in Large Plastic Flow and Fracture with
Emphasis on the Effects of Hydrostatic Pressure, McGraw-Hill, New
1952.
13. W. Johnson and P. B. Mellor, Plasticity for Mechanical Engineers,
Nostrand, Princeton, N.J., 1962.
14. W. Prager, The Theory of Plasticity-A Survey of Recent Ac:hu~ve:me~n
Proc. Inst. Mech. Engrs., London, 169, 41, 1955.
Ch. 2] General References 23
General References
Drucker, D. C., Stress-Strain Relations in the Plastic Range-A Survey of
Theory and Experiment, ONR Rept. NR-D41-D32, 1950.
Goodier, J. N., and P. J. Hodge, Jr., Elasticity and Plasticity, Wiley, New York,
1958.
Hill, R., The Mathematical Theory of Plasticity, Oxford Univ. Press, London,
1950.
Johnson, W., and P. B. Mellor, Plasticity for Mechanical Engineers, Van Nostrand,
London, 1962.
CHAPTER 3
THE STRESS
TENSOR
It is assumed that the reader is familiar with the basic concepts of the theory
of elasticity, including the definitions of stress and strain. However, to avoid
having the student refer to other texts and to refresh the memory of those
readers who have not recently done any work in elasticity, we shall briefly
review in the next three chapters some of these basic concepts, with particular
emphasis on those properties of the stress and strain tensors which are par-
ticularly important in the development of plasticity theory. The reader
thoroughly familiar with elasticity theory may skip directly to Chapter 6.
Stress and strain are second-order tensors, and although we shall not con-
cern ourselves with tensors and their properties as such, it is important that
the student be familiar with the subscript notation known as tensor notation.
This notation is not only a time saver in writing out long formulas or expres-
sions, but it is also extremely useful in derivations and in the proof of
theorems. Furthermore, a major part of the past and present literature on the
subject utilizes tensor notation, and a knowledge of this notation is "'~~''"T''"
in following the literature. In solving specific problems the usual <vu")';u<u,...
notation must, however, always be used. We shall therefore start with a
description of tensor notation.
24
Sec. 3-1] Tensor Notation 25
Similarly, the nine components of the strain tensor are designated by e!i.
Two subscripted quantities are said to be equal if their corresponding com-
ponents are equal. Thus if A!i = B 1h then All = Bll, A 12 = B 12, etc.
If two subscripted quantities are added, their corresponding components
are added. Thus
The advantage here becomes apparent, since we have written down one term
instead of nine. Also the work increment stands out as the scalar product of
the strain increment and the stress. Furthermore, rli de!J represents the work
increment without being tied to any particular system of axes. It would thus
include r 11 de 11 + r 22 de 22 + r 33 de 3s, where these are the principal stresses
and strains.
A system having any number of subscripts is said to be symmetric in two of
these subscripts if the components of the system are unaltered when the two
subscripts are interchanged. Thus a second-order system is called symmetric if
A 11 = -A 11 = 0
Azz = -A22 = 0
A33 = -Ass= 0
P1 = Asz = -A2s
P2 = A13 = -Asl
Ps = A1z = -A21
(3.1.3)
= [~ ~]
It is called the substitution tensor because
(3.1.5)
. b..P
p = 1lm AA (3.2.1)
t.A->0 i l
The important thing to note here is that the unit stress, p, must be referred to
a particular plane. For any other plane passing through the same point, it is
obvious from consideration of Figure 3.2.1 that the force distribution on this
plane, and hence the unit stress, will be different.
The unit stress, p, of course, need not be perpendicular to the plane AB.
In practice, therefore, the stress, p, is decomposed into two components, one
normal to the plane of reference, called the normal stress, and one parallel
to this plane, called the shearing stress. The normal stress is taken as positive
when it is tensile in nature and negative when it is compressive.
To completely specify the stress at a point it is necessary to specify the
stresses at that point on three mutually perpendicular planes passing through
the point. The stress on any arbitrary plane through the point can then be
determined in terms of the stresses on the three perpendicular planes, as will
shortly be shown. Let the three mutually perpendicular planes be the planes
perpendicular to the x, y, and z coordinate axes. Then the stresses acting on
these planes at their point of intersection are as designated in Figure 3.2.2.
The stresses as shown are all positive. The subscripts denote the direction of
the stress. The first subscript designates the normal to the plane under con-
sideration, and the second subscript designates the direction of the stress.
Thus r xy denotes a shearing stress acting on the face of the element that is
perpendicular to the x axis, the stress acting in the direction of the y axis.
As mentioned previously, the normal stress is taken positive when it nrr•rlllt'Po
tension and negative when it produces compression. The positive mrectwtts
of the components of shearing stress on any side of the cubic element
taken as the positive directions of the coordinate axes, if a tensile stress on
same side would have the positive direction of the corresponding axis.
It is seen from the figure that the complete specification of the stress at
point is given by the nine quantities
Sec. 3-ZJ Stress at a Point Z9
OCT 07'
X OT zx YX
-
ax + oy
- +-=
oz -Fx
(3.2.2)
, are the components of the body forces per unit volume. Also
(3.2.4)
The second line of(3.2.4) expresses the fact that the stress tensor is symmetric.
There are therefore in general only six independent components of stress at a
point rather than nine. (Note: There are some peculiar conditions for which
the stress tensor will not be symmetric, as in the case when body moments
act [1 ].)
If we are given the six components of stress at a point with respect to some
coordinate system (x, y, z), we can determine the stresses acting on any plane
through this point. This can be done by consideration of the static equilibrium
of an infinitesimal tetrahedron formed by this plane and the coordinate
planes, as shown in Figure 3.3.1. In this figure we have shown the stresses
acting on the three coordinate planes. These stresses are assumed to be known.
We wish to find the stresses acting on the plane ABC whose normal ON has
direction cosines l, m, and n. Let the area of the infinitesimal triangle ABC be
designated by ~A. Then the areas of the faces AOB, COB, and AOC are equal
to m ~A, l ~A, and n ~A, respectively. Now let the stress vector acting on
the face ABC be designated by Sand its x, y, and z components by Sx, Sy,
and S2 as shown in Figure 3.3.l(b). Then for equilibrium of forces in the
x direction,
or
Similarly,
Sy = lrxy + may + llTzy
S 2 = lrxz + mryz + llUz
z OD ..L ABC
N .I= 00
OA
00
m=-
oc
00
n = 08
c
y
(a)
z
c y
X
(b)
FIGURE 3.3.1 Forces on infinitesimal tetrahedron.
(3.3.3)
Equations (3.3.1) give the x, y, and z components of the stress acting on this
plane, and equations (3.3.3) and (3.3.4) give the normal and shear stresses.
Equations (3.3.1) can also be considered as the boundary conditions that
have to be satisfied by the stress components u 11 at any point on the boundary
of the body. Thus if the element of area ABC is considered to be an element
of the boundary whose normal has the direction cosines /, m, and n, and
Sx, Sy, and S2 are the components of the applied boundary forces at the
point 0, then equations (3.3.1) are precisely the boundary conditions that
must be satisfied by the stress tensor. In tensor notation, if we replace l, m,
and n by /1 , 12 , and /3 , we can write (3.3.1) as
(3.3.5)
Suppose the plane element ABC of Figure 3.3.1 is so oriented that the
resultant stress Son this plane element is normal to the plane; i.e., S = Sn
and Ss = 0. The plane is then called a principal plane at the point, its normal
direction is called a principal direction, and the stress S = Sn is called a
principal stress. At every point in a body there are at least three principal
directions. These principal stresses and principal directions can readily be
found as follows. Assume the element ABC to lie in a principal plane at
point 0 so that S = Sn. Then S has the same direction cosines l, m, and n
as the normal. The components of S in the x, y, and z directions are then
Sx =IS
Sy = mS
Sz = nS
(3.3.9)
where
It can be proved [2] that the cubic equation (3.3.9) has three real roots and
consequently there are (at least) three principal stresses, which will be desig-
nated by a 1 , a 2, and as. Substituting any of these solutions back into equations
(3.3.6) enables one to solve for the corresponding direction cosines l, m, and
n, if in addition the identity !2 + m 2 + n 2 = 1 is used. If the three roots
a 1 , a 2, and as are distinct, the three corresponding principal directions will
be unique and orthogonal. If two of these roots are equal, one direction will
be unique but the other two directions can be any two directions, orthogonal
to the first. If all three roots are equal, there are no unique principal directions
and any three directions can be chosen. This corresponds to a state of hydro-
static stress.
Suppose instead of the axes x, y, and z, a different set of axes, x', y', and z',
were chosen at the point 0. Then the equation for determining the principal
stresses, (3.3.9), would be the same, except that 11 , 12, and 1s would be defined
in terms of the stresses a~, a~, a~, etc., with respect to the new coordinate
ly for 12 and 1s. 11 , 12, and 1s are therefore called the first, second,
third invariants of the stress tensor, respectively. We will show later
the first and second invariants are particularly important in plasticity
34 The Stress Tensor [Ch. 3
/1 = a1 + az + as
l2 = -(ala2 + a2as + aaal) (3.3.11)
Let us take the coordinate axes in the principal directions. Then the shear
stresses referred to these axes are zero and the normal and shear stresses on
some oblique plane, with direction cosines with respect to these axes of!, m,
and n, are, from equations (3.3.3) and (3.3.4),
Sn = + m 2 a 2 + n2 as
l2a1
(3.4.1)
s; = sr + s~ + sg - s~
But, from (3.3.1), with the shear stresses zero,
Therefore,
We already know that on the principal planes the shear stress is a minirnu1n---'
zero. Let us now find the planes for which it is a maximum; i.e., we seek
values of l, m, and n, such that s. as given by equation (3.4.2) is a u"'""w'"'"
In addition to equation (3.4.2), there is a restriction on the direction
/
2
+ m2 + n2 = 1; i.e., only two of them can be independent.
into (3.4.2) n2 = 1 - m 2 - f2, differentiating the resultant equation
Sec. 3-4] Maximum and Octahedral Shear Stresses 35
0 0 ±1 0 ±V1f2 ±V1/2
m 0 ±1 0 ± v'1/2 0
n ±1 0 0 0
The first three columns obviously give the direction cosines of the coordinate
planes which are principal planes, and therefore the shearing stresses on these
planes are zero; i.e., they are a minimum. The three last columns give direc-
tion cosines of 45° angles. These planes, therefore, bisect the angles between
the coordinate axes. On these planes the shearing stresses are maximum.
Designating these stresses by r 1 and substituting these direction cosines into
equation (3.4.2), the values of the shearing stresses are obtained as
rl = ±-t(a2 - aa)
T2 = ±-t(al - aa) (3.4.4)
Ta = ±!(al - a2)
the maximum shearing stress acts on the plane bisecting the ar1gle
the largest and smallest principal stresses and is equal to half the
...~·~"11"" between these principal stresses.
we compute the normal stresses on these planes and designate them by
we get, from (3.4.1),
N1 = !(a2 + aa)
N2 = !(al + aa) (3.4.5)
Na = !h + a2)
36 The Stress Tensor [Ch. 3
so that the normal stress on each of these planes is equal to the average of the
principal stresses on the two planes whose angle it bisects.
If the plane ABC of Figure 3.3.1 is so oriented that OA = OB = OC, then
the normal ON will make equal angles with all the axes, and
1
l=m=n= ±v'"3 (3.4.6)
(3.4.7)
That is, the normal stress on this plane is equal to the mean stress, and the
shear stress is
In terms of the stress invariants the octahedral shear stress can be written
'Tact=
v'2 (J21 + 312)1/2
3
Sec. 3-5] Mohr's Diagram 37
which gives the octahedral shear stress in terms of the stress components
referred to an arbitrary set of axes.
The result of the previous section can be obtained graphically by the use
of a diagram proposed by Mohr [3]. This approach is particularly useful in
discussing a pair of parameters called Lode's variables which are frequently
encountered in plasticity theory. A full discussion of Lode's variables will be
postponed until later. Mohr's diagram can be obtained in the following way.
Let us take the principal directions for the coordinate axes and consider the
normal and shear stresses acting on an arbitrary plane oblique to these axes.
Instead of Sn and Ss we will for convenience use a and, r for the normal and
shear stresses on this plane. Then in terms of the principal stresses equations
(3.4.1) and (3.4.2) become
f2 = -r
2
+ (a - a2)(a - as)
(a 1 - a2 )(a1 - as)
m 2 = -r 2 + (a - as)(a - a 1) (3.5.2)
(a 2 - as)(a 2 - a 1 )
n2 = -r
2
+ (a - a 1)(a - a 2)
(as - al)(as - a2)
(3.5.3)
(3.5.4)
or
or (3.5.6)
which represents the region exterior to the circle C3 with center at (a1 + a 2 )/2
and crossing the a axis at a 1 and a 2 •
It follows therefore from (3.5.4), (3.5.5), and (3.5.6) that the admissible
values of rand a lie in the shaded region shown in Figure 3.5.1 bounded by
the circles C1 , C2 , and C3 • The maximum shearing stress, as is clear from the
figure, is represented by the largest ordinate, AB, which is the radius of the
circle C2 and is therefore equal to (a 1 - a 3 )/2. To determine the orientation
of the plane that has this shearing stress, we make use of equation (3.5.2).
The value of a corresponding to the maximum shearing stress is equal to
(a1 + a 3)/2 (the center of the circle C2 ). Substituting these values of rand a
into equations (3.5.2), we get
We have thus obtained the same values for the maximum shearing stress as
in the previous section.
It is convenient in plasticity theory to split the stress tensor into two parts,
one called the spherical stress tensor and the other the stress deviator tensor .
.The spherical stress tensor is the tensor whose elements are am oli. where am
the mean stress, i.e.,
(3.6.1)
(3.6.2)
'T'xy (3.6.3)
0
am
u,~J
2a 1 - a2 - a3
0 0
2
2a 2 - a3 - a1
0 0
3
2a 3 - a1 - a 2
0 0
3
where
Jl = 0
'2 = ten + 3/2) (3.6.6)
J3 = -i-7(2Jr + 9/1/2 + 27/3)
One advantage of using the stress deviator tensor is now apparent. The
first invariant of this tensor is always zero. This can also be seen by taking
the sum of the diagonal elements in equation (3.6.3) or (3.6.4).
The invariants 1 2 and J 3 can, of course, be written in terms of the stress
components a 11 • For example,
(3.6.7)
(3.6.10)
PURE SHEAR
a3 =0 (3. 7.1)
42 The Stress Tensor [Ch. 3
It can readily be shown [4] that a necessary and sufficient condition for a
state of pure shear to exist is
(3.7.2)
Problems
1. Write out the expression for the work increm.-nt
dW = T!J defJ
2. Given the tensor TucTicJ·
(a) What is the order of this tensor?
(b) If i, j, and k range from 1 to 3, how many components does this tensor
have?
(c) Write out the components.
3. The general Hooke's law is given by
References
L I. S. Sokolnikoff, Mathematical Theory of Elasticity, 2nd ed., McGraw-Hill,
New York, 1956, p. 42.
2. Ibid., p. 17.
3. 0. Mohr, Abhandlungen aus dem Gebiet der technischen Mechanik, 2nd ed.,
Wilhelm Ernst und Sohn, Berlin, 1914, p. 192.
4. C. E. Pearson, Theoretical Elasticity, Harvard University Press, Cambridge,
1959, p. 57.
General References
Sokolnikoff, I. S., Mathematical Theory of Elasticity, 2nd ed., McGraw-Hill,
New York, 1956.
and J. N. Goodier, Theory of Elasticity, McGraw-Hill, New
CHAPTER 4
THE STRAIN
TENSOR
44
Sec. 4-1] Strain at a Point 45
Pt(x,y,z)---.
--- -- --
A
Uo =X~- Xo
Vo = Y~- Yo (4.1.1)
w0 = z~ - z0
u = x'- x
v = y'- y (4.1.2)
w = z'- z
8u ou ou
U = Uo + ox Ax + oy Ay + oz Az
(4.1.3)
The Strain Tensor [Ch. 4
46
where higher-order terms are neglected, since P is taken in the neighborhood
of P • Making use of (4.1.1) and (4.1.2),
0
, ) ( , ) 8u A A 8u A
( X - X - Xo - Xo = -ox x + -8u
By Y + -8z
2
(4.1.4)
, ) ( , ) 8v A 8v A Bv A
( Y - Y - Yo-Yo=-
ox x +-
By +-
YBz
2
A ow A ow A
(
I )
z - z -
( I
Zo - Zo
)
= -ow
ox x + -By Y + -8z z
Now the change in the components of the vectors A and A' can be written
8u 8u 8u
8u1 8u1 8u1 By 8z
ox
8x1 8x 2 8x 3
8v 8v 8v
8u 2 8u 2 8u 2
ox By 8z
Ut,J = Bx 1 8x 2 8x3
ow ow ow
8u 3 8u3 8u 3
ox By 8z
8x 1 8x 2 8x 3
Sec. 4-1] Strain at a Point 47
or (4.1.8)
equation (4.1.9) must be true for all values of Ax, Ay, and A 2 , a necessary
sufficient condition that the transformation (4.1.6) represent a rigid body
is
ou = ov =ow=
ox oy oz 0
(4.1.10)
ou ov ou ow
oy = -ox oz -ox
48 The Strain Tensor [Ch. 4
That is, for rigid body motion the tensor u1,1 of equation (4.1.7) is skew-
symmetric.
Now every second-order tensor can be decomposed into a symmetric
tensor and a skew-symmetric tensor in one and only one way (see Problem 4,
Chapter 3). It follows therefore that if we decompose the tensor u1,1 into
symmetric and skew-symmetric parts, the skew-symmetric part will represent
rigid body motions, whereas the symmetric part will represent pure deforma-
tion. Therefore, let
Ut,J = -!(ut,J + UJ,t) + !(Ut,J - UJ,t) (4.1.11)
or
Ut,f = 8 tJ + Wtj
where
ou .1(8u + ov) .1(8u + ow)
ox 2
oy ox oz ox
2
I cos e= A'. Jj' = oAx oBx + (Ay + oAy) oBy + (Bz + oBz) oAz
;:;; Ay oBy + Bz oAz (4.2.3)
.ii'. B'
p'''
f..(-8Az
0 p y
8A 2 = e 2 yAy
(4.2.5)
8By = By 2 B 2
Also cos 8 = cos (7r/2 - a), where a is the decrease in the right angle between
A and B. Or
cos 8 = sin a ~ a
a = Byz + By =
2 2ey 2
Hence eyz represents half the decrease in the right angle between two vectors
that were initially directed along the positive y and z axes. Also, from Figure
4.2.2 and equation (4.2.5),
SA
L POP' ~ tan POP' = Ayz = Bzy
Thus angles POP' and ROR' are equal. If we now rotate the
R 'OP' Q' through an angle ey 2 about the origin, we obtain a figure such
that of Figure 4.2.3. This represents a sliding or shear ofthe elements
to the xy plane, the amount of the sliding being proportional to the
of the element from the xy plane. Hence eyz is called a shear strain.
derivations can be made for exy and exz·
Sec. 4-3] Finite Deformations 51
R'
y
FIGURE 4.2.3 Shear deformation.
(4.3.1)
(4.3.2)
a1 = a1(x1, x2, x3 )
(4.3.3)
da 1 = a1,1 dx1 = a1," dx"
52 The Strain Tensor [Ch. 4
Therefore,
Therefore,
(4.3.4)
U1 = X1 - al
a1 = X1 - U1
or
Hence
The tensor
(4.3.6)
is called the Eulerian strain tensor. Some typical terms in engineering notation
are
We see that if the derivatives are small, so that their products can be neglected,
these reduce to the same expressions as previously obtained for the infini-
tesimal strains. The simple physical interpretations given in Section 4.2 for
small strains are, however, no longer applicable in this case. Thus if E 1 is the
change in length per unit length of an element finally parallel to the x axis,
it can be shown [2) that
Sec. 4-4] Principal Strains. Strain Invariants 53
In the study of the stress at a point we found that there exist at least three
mutually orthogonal planes which have no shear stress acting on them, i.e.,
the principal planes. We now ask ourselves, do there exist in a similar fashion
planes with no shoor strain? By this we mean planes whose normals will not
change orientation when the body is strained. Thus a vector A, which was
originally normal to such a plane, will either shorten or lengthen but will not
change direction. The answer to this question is yes. Such planes, as in the
case of stress, are called principal planes, the normal directions to these planes
are the principal directions, and the corresponding strains are called principal
strains. To find these directions and the corresponding strains we proceed as
follows.
Consider a vector A normal to the plane ABC as shown in Figure 4.4.1.
Upon straining it is assumed that A changes length by an amount BA but its
direction remains the same; i.e., BA is in the same direction as A. Since A
and BA are parallel, the components of A and of BA are proportional; i.e.,
(4.4.1)
BA (4.4.2)
e=A
54
The Strain Tensor [Ch. 4
l
z
X
FIGURE 4.4.1 Principal strain vector.
To illustrate again the power of the tensor notation, equations (4.4.3) can
be written
or
Byz =0 (4.4.6)
Equation (4.4.6) is of exactly the same form as (3.3.8) with stresses replaced
by strains. All the remarks and derivations made there, therefore, apply here
just as well. Thus (4.4.6) will have three real roots el> e2 , and e3 , corresponding
to the three principal strains. The invariants appearing in the cubic equation
(4.4.7)
are
I~ = Bx + By + B 2
+ e 2 + e3
I{ = e 1
If = -(e 1 e 2 + e 2 e 3 + e 3 e 1 ) (4.4.9)
I~ = e 1e 2 e 3
are designated by 1, 2, and 3. Let the vector A suffer a small strain so that the
point P moves to Q. The strain is thus composed of two parts: a linear strain
of amount e = RQJA, and a rotation or shear of amount 8 = RPJA, for
small 8. To calculate the linear strain, e, we have
2
(A + RQ)2 = (A1 + 8A1) 2 + (A2 + 8A2)2 + (A3 + 8A3) (4.5.1)
A 2 + 2A(RQ) ;;:::: Ai + A~ + A~ + 2(A 1 8A 1 + A2 8A 2 + A3 8A3)
(4.5.2)
From the basic equation (4.1.14), which is given in expanded form in (4.4.4),
we have
8A1 = e1A1
8A 2 = e2A2
8A3 = e3A3
Equation (4.5.4) gives the linear strain for a vector having direction cosines
!, m, and n with respect to the principal strain directions, in terms of the
principal strains. To determine the shear 8 = RP/A, we have
2 2 2 2
2 +
e (PQ)
82 _ - --
A2
_ - 1-
A
) +-
A
(SA
2) + -3)
A
(SA (SA
or (4.5.5)
or (4.5.6)
Comparing equations (4.5.6) with (3.4.2), it is seen that they are of identical
forms, with s. replaced by 8 and the principal stresses by the principal
strains. The maximum shear strains and the corresponding directions can
therefore be obtained in the exact same way as for the stresses. Thus desig-
nating the maximum shear strains by Y1> y 2 , and y 3 , we can write directly, by
analogy to (3.4.4),
Y1 = ±!(e2 - ea)
Y2 = ±!(el - ea) (4.5.7)
Ya = ±!(el - e2)
and the same table of direction cosines is applicable as given on page 35.
Thus the maximum shearing strain acts on a plane bisecting the angle between
the maximum and minimum principal strain directions and is equal to half
the difference of these strains.
If we consider the octahedral planes for which I = m = n = ± ljv3, we
see from (4.5.6) that the shear strain on these planes, which we designate
Yoct> is given by
As in the case of the stress tensor, the strain tensor can be separated into
two parts, a spherical part q11 and deviator part e11 • The spherical part is
given by
(4.6.1)
where Bm = t(e1 + Bz + B3) is the mean strain. The deviator strain then
becomes
2Bx- B2 - By
Bxy Bxz
3
2By- Bz- Bx
Bxy Byz (4.6.2)
3
2B 2 - B x - By
Bxz Byz
3
2B 1 - Bz - B3
0 0
3
2B 2 - B3 - B1
el1 = 0
3
0
2B3 - B1 - Bz
0 0
3
or abc + abc(e 1 + e2 + B3 )
Sec. 4-7] Compatibility of Strain 59
neglecting products of the strains. The change in volume per unit volume is
then
The spherical strain tensor is thus proportional to the volume change. The
deviator strain tensor then represents a pure distortional strain.
By analogy with the stress deviator tensor discussed in Section 3.6, the
invariants of the strain deviator tensor are
J{ = 0
J~ = -!(I? + 3!~) (4.6.4)
J~ = -f?- (2!? + 9I{I~ + 27!~)
or
J~ = ifh - e2) 2 + (e2 - ea) 2 + (ea - e1) 2]
= i[(ex - ey) 2 + (ey - e2 )2 + (ez - ex) 2 + 6(e~y + e;z + e~x)]
= -(e1e2 + e2es + e3 e1 ) (4.6.5)
J~ = e1e2ea
where
Also it follows that
(4.6.6)
The strain tensor eu defines six strain components in terms of three dis-
placements; i.e.,
8u 8v
e =- e =-
x ox Y By
(4.7.1)
_(ou ov)
1
exy- 2 8y + OX eyz =
1
2
(8vf}z + ow)
By
_
exz-
1
2
(8u ow)
f}z + ox
solution will not be unique, for the strains represent pure deformation,
whereas the displacements include rigid body motion which have no effect
on the strains. The problem can, however, be made unique by specifying the
rigid body motion, i.e., specifying the displacement and rotation at some
point in the body.
However, a more difficult problem is encountered in calculating the dis-
placements from the strains. Equations (4.7.1) are six equations for the three
unknowns u, v, and w. It is evident, therefore, that these equations will not
have a solution for any arbitrarily chosen strains, but that some restrictions
must be placed on the strains in order that equations (4.7.1) have a solution.
This can also be seen from the following physical considerations. Assume
the body is divided into infinitesimal cubes. Let all these cubes be separated
from each other and let each of the cubes be subjected to some arbitrary
strains. It is obvious that if we now try to put all the cubes together, we will,
in general, no longer be able to fit all the cubes together the way they were
before, to produce a continuous body. Between some of the cube boundaries
there will be gaps; others will overlap. This shows that there must be some
relationships between the strains at the different points of a body in order
that the body remain continuous after straining, i.e., that the displacements
be continuous functions of the coordinates. These relationships are called
the compatibility relations, or, sometimes, the continuity relations.
The compatibility equations are found in all standard texts on elasticity.
They are derived, for example, in a particularly elegant fashion in reference
[3]. In engineering notation these equations are
(4.7.2)
8 ( 8eyz
8ezx
- --
ox - +8exy)
8x + oy -
8z
Equations (4. 7.2) are necessary and sufficient conditions that the strain
components give single-valued displacements for a simply connected region.
For a multiply connected region, however, these conditions are necessary but
generally not sufficient.
Just as for the stress equations of Chapter 3, it should also be emphasized
here that all the relations presented in this chapter, including the com-
patibility relations (4.7.2), are independent of the material properties and
therefore hold for both elastically and plastically behaving materials.
The shear strains defined herein differ from the shear strains defined by
many authors by a factor of 1/2. Thus Timoshenko, for example, defines the
shear strains by
ou ov
YxY = oy + ox
ou ow
Yxz = OZ + OX
ov ow
YYz = oz + oy
The factor 1/2 used herein is necessary for the tensor definition of strain as
exemplified by equation (4.1.12). Care must be exercised in reading the
literature to determine which definition of shear strain is used.
Problems
1. Obtain the equations representing rigid body displacements; i.e., show that
U = a + WyZ - WzY
V = b +W X
2 - WxZ
W = C + WxY - WyX
2. Let the rectangle of Figure 4.2.3 be a square. Show that the shear strain is
equal to the extension of the diagonal of this square.
3. The displacement vector at a point in a body is given by
-0.005 -0.004
-0.~04
e11 =
[
0.001
0 o.Ll
Determine:
(a) The principal strains.
(b) The direction cosines of the principal directions.
(c) The largest shearing strain.
(d) The octahedral shear strain.
5. If a rectangular parallelepiped with initial dimensions of 1 in. in the x direc-
tion, 2 in. in the y direction, and 3 in. in the z direction is strained to the
condition of Problem 4, determine the final dimensions.
6. Derive the Lagrangian strain tensor.
7. Show that in plane polar coordinates the infinitesimal strains are given by
ou 1 ov u
e, = or 89
= ro(J + r
ov 1 ou v
8 9
' = or + o(J - r r
where u and v are the displacements in the radial and tangential directions,
respectively.
8. Assuming axial symmetry, derive the compatibility equation in plane polar
coordinates.
9 Verify equations (4.5.9) and (4.5.10).
10. Verify equations (4.6.4) and (4.6.5).
References
1. I. S. Sokolnikoff, Mathematical Theory of Elasticity, McGraw-Hill, New
York, 1956, p. 21.
2. Ibid., p. 31.
3. Ibid., p. 25.
4. W. Prager and P. G. Hodge, Theory of Perfectly Plastic Solids, Wiley, New
York, 1951, p. 118.
References 63
General References
Novozhilov, V. V., Theory of Elasticity, Pergamon Press, London, 1961.
Sokolnikoff, I. S., Mathematical Theory of Elasticity, McGraw-Hill, New York,
1956.
Timoshenko, S., and J. N. Goodier, Theory of Elasticity, McGraw-Hill, New
York, 1951.
CHAPTER 5
ELASTIC
STRESS-STRAIN
RELATIONS
Hooke first proposed a linear relation between stress and strain for a load
applied in one direction. The generalization of Hooke's law to three dimen·
sions is given in Problem 3, Chapter 3. For an isotropic material this becomes,
using the tensor subscript notation and including thermal strains,
E
G = 2(1 +f.") (5.1.2)
and
(5.1.3)
(5.1.4)
where
8= Bx + By + B = 2 If
(5.1.5)
.\ = 1-"E
(1 + f.")(l - 2fl,)
1
By = E [ay - f."( ax +a 2 )] + aT
(5.1.6)
1 1 + 1-"
Byz = 2G 7'yz = ~ 7'yz
1 1 + 1-"
Bzx = 2G,7'zx = ~ 1'zx
1- 2fl,
8= - E - 0 + 3aT (5.1.7)
1 - 2fl,
Bm = --E-am +aT (5.1.8)
66 Elastic Stress-Strain Relations (Ch. 5
where em and am are the mean strain and mean stress, respectively. Finally,
r
combining (5.1.8) and (5.1.1) results in
(5.1.9)
where e!J and Sli are the strain deviator and stress deviator tensors, respec-
tively. Thus the deviators of the stress and strain tensors are related to each
other, in the elastic case, by the simple equation (5.1.9), whereas the spherical
stress components are related to the spherical strain components by equation
(5.1.8).
It also follows that
JJ. = Jl = 0
(5.1.10)
J~ = 4~2J2
2 1 2
(5.1.11)
Yoct = 4G2 '~"oot
Let us now denote the strains due to temperature rise by primes and
strains due to stresses by double primes. Then assumptions 1 and 2
us to write
e~ = e~ = e~ = aT
Sec. 5-2] Elastic Strain Energy Functions 67
and now making use of assumption 3, the total strains are as given by the
first three of equations (5.1.6), whereas Hooke's law and assumption 4 give
the last three of equations (5.1.6). Note that assumption 2 may be modified
to the extent that a may be a function of temperature.
(5.2.1)
U = Gelie11
A
+Z 8 2
-
2G + 3A aT8 (5.2.2)
2
Making use of the definitions of the invariants of the strain and strain
deviator tensors, equation (5.2.2) can be written after some algebraic manip-
ulations as
U=
20
+ 3A (I~
6
2
- 3aTI~) + 2GJ~ (5.2.3)
in changing the volume; the second term is the energy of distortion, which is
designated by Ua and can be written
Ua = 2GJ~ = 3Gy~ot
or, from (5.1.10),
Ud = 1G J.2 = 4G
3 2
Toot (5.2.4)
2
To solve an elastic problem we have to find the stresses and the strains
which will satisfy the previously derived equations. Thus the stresses must
satisfy the equilibrium equations, (3.2.2) and (3.2.3), as well as the boundary
conditions (3. 3. 5). The strains must satisfy the compatibility equations (4. 7.2).
Finally, the stresses must be related to the strains through the stress-strain
relations (5.1.1) or their equivalent. The problem of finding a set of stresses
and strains satisfying the above relations is known as the .first boundary-value
problem of elasticity. Alternatively, it is possible to reduce the above set of
equations to three equations, called the Navier equations, involving only the
displacement u1 [2].
and the plastic problem occurs in the stress-strain relations. If plastic flow
occurs, the linear generalized Hooke's law no longer holds and the relations
between stress and strain become nonlinear in a very complicated manner.
What happens to the relations between stress and strain when plastic flow
occurs will be the subject of Chapters 6 and 7.
Problems
1. Show how equation (5.1.4) can be obtained from equation (5.1.1).
2. Derive equation (5.1.9).
3. Given the following stress tensor at a point:
10,000 1,000 -8,000
1,000 -6,000 6,000
-8,000 6,000 20,000
Assume the material behaves elastically with an elastic modulus of 30 x 106
and Poisson's ratio of 0.3. Determine the strain deviator tensor. Calculate
the mean strain assuming the temperature rise at the point is 300°F and that
a = 7.5 x 10- 6 in./in.;aF.
4. Verify equations (5.1.10) and (5.1.11).
5. Prove that the axes of principal stress coincide with the axes of principal
strain for an isotropic material obeying Hooke's law.
6. Prove that for the material of Problem 5, Mohr's stress diagram will be
similar to Mohr's strain diagram.
7. Derive equation (5.2.3).
References
1. I. S. Sokolnikoff, Mathematical Theory of Elasticity, McGraw-Hill, New
York, 1956, p. 84.
2. Ibid., p. 73.
General References
Hoffman, 0., and G. Sachs, Introduction to the Theory of Plasticity for Engineers,
McGraw-Hill, New York, 1953.
Sokolnikoff, I. S., Mathematical Theory of Elasticity, McGraw-Hill, New York,
1956.
CHAPTER 6
CRITERIA FOR
YIELDING
70
Sec. 6-2] Examples of Yield Criteria 71
different principal directions. The question here is: For what combinations
of loads will the cylinder begin to yield plastically?
Another simple example is the plane stress problem of a thin rotating disk
with or without temperature gradients. At every point of the disk (except
possibly at the rim) there exists a state of biaxial stress. The question again
arises: For which states of biaxial stress will the disk deform plastically? The
criteria for deciding which combination of multiaxial stresses will cause
yielding are called yield criteria. The first step of any plastic flow analysis is
to decide on a yield criterion. The next step is to decide how to describe the
behavior of the material after yielding has started. In this chapter we shall
discuss the choice of a yield criterion.
Numerous criteria have been proposed for the yielding of solids, going as
far back as Coulomb in 1773. Many of these were originally suggested as
criteria for failure of brittle materials and were later adopted as yield criteria
for ductile materials. Some of the more common ones will be briefly discussed.
Although some of these theories are no longer in use, they are included here
both for their historic interest and to give the reader a feeling for the type of
approach used in promulgating yield criteria.
This theory assumes that yielding occurs when one of the principal stresses
becomes equal to the yield stress in simple tension a 0 , or the yield stress in
compression ao,c· Thus if a 1 is the maximum principal stress and a 2 is the mini-
mum principal stress, yielding will occur in tension when a 1 = a 0 and it will
occur in compression when a 2 = a 0 , 0 • For a material with the same yield in
tension and compression, this criterion becomes
or (6.2.1)
A simple plot illustrating this criterion for the case of biaxial stress with
aa = 0 is shown in Figure 6.2.1. The coordinates are the remaining principal
stresses a 1 and a 2 • Yielding occurs when the state of stress is on the boundary
of the rectangle, for then one of the stresses is at the yield point in tension or
compression. For example, consider a thin-walled cylinder subjected to an
72 Criteria fo:r Yielding [Ch. 6
0"2
uo
o,c /
....1/-' cro
CTj
cro,c
This theory assumes yielding will occur when the maximum value of the
principal strain equals the value of the yield strain in simple tension (or
compression), e0 = a 0 /E. Thus if e 1 is assumed to be the largest strain in
absolute value, yielding will occur when
(6.2.2)
crz
This theory (sometimes called the Coulomb theory) assumes that yielding
will occur when the maximum shear stress reaches the value of the maximum
shear stress occurring under simple tension. The maximum shear stress is
given by equation (3.4.4) and is equal to half the difference between the
maximum and minimum principal stresses. For simple tension, therefore,
since a 2 = a 3 = 0, the maximum shear stress at yield is !a0 • The Tresca
criterion then asserts that yielding will occur when any one of the following
six conditions is reached:
(6.2.4)
if a1 > 0, a2 < 0
if a1 < 0, a2 > 0
if a 2 > a1 > 0
(6.2.5)
if a 1 > a 2 > 0
if a 1 < a 2 < 0
if a 2 < a1 < 0
74 Criteria for Yielding [Ch. 6
A plot in the a 1a 2 plane for this yield criterion is shown in Figure 6.2.3. It is
to be noted that one limitation of this theory is the requirement that the yield
or
That is, the yield stress in pure shear is t the yield stress in simple tension.
This theory assumes that yielding will occur when the total strain
per unit volume equals the total strain energy per unit volume at yielding
uniaxial tension or compression. The total strain energy at yield in
tensile test is
Sec. 6-2] Examples of Yield Criteria 75
U= -! (a1e1 + a 2 s 2 + asss)
= 2~ [a~ + a~ + a~ - 2p,(a 1 a 2 + a 2 as + asa1)]
(6.2.6)
(6.2.7)
The distortion energy theory (also associated with Hencky) assumes that
yielding begins when the distortion energy equals the distortion energy at
yield in simple tension. Thus from equation (5.2.4),
_lr_3 2
Ud - 2G"2- Toot
40
(6.2.8)
This plots as an ellipse, called the von Mises ellipse, in the a 1 a 2 plane, as
shown in Figure 6.2.4. For the case of pure shear
and the von Mises criterion would predict yielding to occur when
or k= ~3
That is, the yield stress in pure shear is l/V3 times the yield stress in simple
tension. Thus the von Mises criterion predicts a pure shear yield stress which
is about 15 per cent higher than predicted by the Tresca criterion. It will
subsequently be shown that this is the maximum difference between the two
criteria.
The von Mises yield criterion usually fits (but not always) the experimental
data better than the other theories, and it is usually easier to apply than
Tresca criterion because no knowledge is needed regarding the
magnitudes of the principal stresses. For these reasons, this criterion
widely used at the present time. If, however, the relative magnitudes of
principal stresses are known, as, for example, in the case of the
tube discussed in Chapter 8, the Tresca criterion is easier to apply.
Sec. 6-2] Examples of Yield Criteria 77
Von Mises originally proposed his criterion because of mathematical
convenience. Hencky later showed that it was equivalent to assuming that
yielding will take place when the distortion or shear strain energy reaches a
critical value, as shown above. Also, since the octahedral shear stress is
equal to
'roct,O =
v2
3 Go
Toot = Toot,O
That is, yielding will occur when the octahedral shear stress reaches the
octahedral shear stress at yield in simple tension.
Alternatively, the criterion (6.2.8) can be looked upon as stating that
yielding will occur when the second invariant J 2 of the stress deviator tensor
reaches a critical value, i.e., the value of J 2 at yield in simple tension. The
assumption that the yield criterion should depend on the invariants of the
stress deviator tensor is generally accepted, as will be discussed in the next
section.
Mohr extended the maximum shear stress theory by assuming that the
critical shear stress is not necessarily equal to the maximum shear stress but
depends also on the normal stress acting on the shearing plane. In general,
the greater the normal stress, the lower is the critical shear stress. Mohr's
theory therefore takes into account the effect of mean stress, which has
experimentally been shown to be important in some cases, particularly with
regard to fracture. Mohr's theory can also take into account differences in the
points in tension and compression, as exemplified by the internal friction
This is a special case of Mohr's theory, in which the critical shear stress is
ouu"'""'u to be a linear function of the normal stress acting on the plane of
Criteria for Yielding [Ch. 6
78
maximum shear. This results in the following conditions for the biaxial case
[la]: If a 1 > a 2 > a 3 , a 2 = 0,
ao (6.2.10)
a l - - a3 = ao
ao,c
2
a3
2
where a ,c is the yield stress in compression. If a 1 > > a , a = 0, then
0
(6.2.11)
If a 1 > a 3 , both negative, then
a3 = -ao,c
cr1 =- ao,c
The various yield criteria listed by no means exhaust the available list.
Besides, there are other, more complicated theories which attempt to take
into account anisotropy. For example, a yield criterion proposed by
[2] for anisotropic materials with different yields in tension and compression
has the form (for biaxial stress)
Sec. 6-3] Yield Surface. Haigh-Westergaard Stress Space 79
where a 11 and a 12 are the tensile yields in the a 1 and a 2 directions, a 01 is the
compressive yield in the a 1 direction, and To is the shear yield strength.
In Section 6.2 we discussed several yield criteria and also plotted several
two-dimensional plots for biaxial stress cases, showing the curves at which
yielding takes place. In the most general case, the yield criterion will depend
on the complete state of stress at the point under consideration and will
therefore be a function of the nine components of stress at the point. Since
the stress tensor is symmetric, we can reduce this function to a function of the
six independent components of the stress tensor. This yield criterion for a
virgin material is then essentially the extension of the single yield point of the
uniaxial tensile test to the six-component stress tensor. For a material loaded
to the initial yield, it can be expressed by the relationship
F(alf) = K (6.3.1)
(6.3.2)
(6.3.3)
80 Criteria for Yielding [Ch. 6
(6.3.4)
J1 = s1 + s2 + s3 = o
J2 = -!-(S~ + S~ + S5) = -(S1S2 + S2S3 + S3Sl) (6.3.5)
J3 = tcs~ + s~ + s~) = s1s2sa
we can write
(6.3.6)
Subject therefore to the above two assumptions, the yield criterion has been
reduced to a function of the two nonzero invariants of the stress deviator
tensor. We note in passing thatj(J2 , J 3 ) is symmetric in the principal stresses,
which is to be expected since in an isotropic material all of the principal
stresses must play the same role in yielding. Thus whatever yield function is
chosen, it must be symmetric in the principal stresses.
The two most widely used yield criteria, the Tresca maximum shear
criterion and the von Mises yield criterion, discussed in Section 6.2, are
specific cases of (6.3.6). The von Mises criterion is by far the simplest one
that can be associated with equation (6.3.6),
(6.3.7)
and we see the great complexity of the Tresca criterion in its general form
compared to the von Mises criterion. Only in the case, as previously discussed,
where the maximum and minimum principal stresses are known a priori can
the Tresca criterion be reduced to the simple form
(
Sec. 6-3] Yield Surface. Haigh-Westergaard Stress Space 81
We can learn a great deal about the yield surface defined by equation
(6.3.3) from simple geometric considerations. We introduce the (a 1 , a 2, as)
coordinate system, which represents a stress space called the Haigh- Wester-
gaard stress space [3]. Every point in this space having coordinates al> a 2 , and
as is a possible stress state. Consider a line ON as shown in Figure 6.3.1
passing through the origin, and having equal angles with the coordinate axes.
Then for every point on this line the stress state is one for which
(6.3.10)
(6.3.11)
where p is the distance along the normal from the origin to the plane. Hence
the spherical component of the stress tensor increases linearly with the
distance of the plane from the origin. On the plane passing through the
origin, the spherical stress is zero, the equation being a 1 + a 2 + as = 0. This
plane is called the 1r plane.
Now consider any arbitrary stress state such as at point P with stress
:;comr>onrents a 1 , a 2 , and as. The stress vector OP can be decomposed into two
cornp<ments, the component A parallel to ON and the component 1i per-
pertdli~Ul;ar to ON, the latter parallel to the 1r plane. The component A can
82 Criteria for Yielding [Ch. 6
1 1 1
A = VJ a 1 + yj a2 + yj a 3
or A= v3 am (6.3.12)
To determine Jj we have
= ar + a~ + a~ - 3a~
= (al - am)2 + (a2 - am)2 + (aa - am)2
= sr + s~ + s~
= 212 (6.3.13)
equal angles with each other, 120°, as shown. It can be shown from symmetry
considerations that the yield locus must have the same shape in each of the
twelve 30° sectors dividing the 1r plane as shown in the figure. This follows
from the following considerations. The yield surface must be symmetric in
the principal stresses since it certainly does not matter, for example, if we
interchange the values of a 2 and as. It follows, therefore, that the lines
bisecting the angle between any two principal axes in the plane must be lines
of symmetry. The lines a 1 , - a 1 ; a 2 , - a 2 ; and as, -as are therefore lines of
symmetry and we now have six symmetric sectors. Furthermore, if we assume
equal yielding in tension and compression, then if we go from a point Q on
the yield locus to the point Q', where all the stresses have reversed signs, we
should again be on the yield locus. Therefore, if a 1 goes to -a1 , we must
have symmetry about a line perpendicular to the a 1 axis; if a 2 goes to - a 2 ,
we must have symmetry about a line perpendicular to the a 2 axis; and if as
goes to -as, we must have symmetry about the line perpendicular to the as
axis. We have thus divided the yield locus into 12 symmetric sectors, each of
, and we need only consider the stress states lying in one of these sectors.
To determine the location on the 1r plane of the projection of the point P
on the yield cylinder, we proceed as follows. Let r and 8 be the polar
~·~··"""""' of the point Pin the 1r plane, 8 being measured from the horizon·
axis, and let a and b be the horizontal and vertical components of r, as
in Figure 6.3.3. P has components a 1 , a 2 , and as, and in the 1r plane
will project as V2/3 a 1 , V2/3 a 2 , and V2/3 as, respectively. This can
seen in Figure 6.3.4, which shows the plane containing the vector a 1 the
Criteria £or Yielding [Ch. 6
84
normal to the plane, ON, and the projection of a 1 on the 7T plane. It is obvious
1
from the figure that the projection of a 1 on the 7T plane is equal to V2{3 a •
From Figure 6.3.3:
a = V2{3 a2 cos 30° - V2/3 a 1 cos 30° = (a 2 - lll)fVZ
b = V2/3 as - V2f3 a 2 sin 30° - V2f3
a 1 sin 30°
= l/v'6 (2as - a 2 - a1) (6.3.14)
rz = az + bz =
(a 2 -a
2 1
)2 + l(2as - az - a1)2
2
= i-[(al - az) 2 + (az - as) + (as - a1)2]
2
= [(al - am) 2 + (az - am) 2 + (as - am) ]
= 21 (6.3.15)
2
1T plane
Let us now consider the yield locus for the von Mises yield criterion
[equation (6.3.7)]:
then be a circular cylinder whose axis is the line ON equally inclined to the
stress axes.
The yield locus for the Tresca maximum shear criterion is a regular hexagon
inscribed in the von Mises circle, as shown in Figure 6.3.5. This can be proved
as follows.
Consider the sector between the a 2 and -a1 axes in Figure 6.3.5. For any
stress state in this sector,
The Tresca yield condition for this sector is then a 2 - a 1 = a 0 • This will be
represented by a line parallel to the a 3 axis, since it is independent of a 3 , the
distance of this line from the a 3 axis being equal to l/v2 (a2 - a 1 ) =
1/Vl a0 [by equation (6.3.14)]. In a similar manner, a straight line is obtained
for each sector, thus forming the hexagon shown in the figure. The corners of
hexagon will touch the von Mises circle, as can be seen again from the
sector, since the radius to the corner is equal to
r = (1/V2)a0 = yi3 a 0
cos 30°
86 Criteria for Yielding [Ch. 6
which is the radius of the von Mises circle. In the stress space the Tresca yield
surface is a regular hexagonal cylinder inscribed in the von Mises cylinder, as
shown in Figure 6.3.6.
It is now evident from Figure 6.3.5 that the maximum difference between
the von Mises and Tresca criteria occurs at 8 = 0° and is equal to
as previously mentioned.
The line at 8 = oo corresponds to pure shear stress states. This follows
from (6.3.15), for e = 0:
2_a--"3'---a-=2'---a-=1 = 1
Uz- a1
or
It can now readily be shown that if the yield locus is assumed to be convex
and one circumscribes the von Mises circle by a regular hexagon, then all
possible yield loci must lie between the two regular hexagons inscribed in,
and circumscribing, the von Mises circle. By the locus being convex, we mean
simply that any straight line in the 7T plane may pierce the locus in at most
two points. It will be proved later that the yield surface must indeed be
convex; for the present we shall merely assume it. Let the point A on the a 3
axis in Figure 6.3. 7 lie on the yield locus. By the symmetry conditions pre-
viously discussed, the points B and C must also lie on the yield locus. The
curve CAB is a convex piecewise smooth curve passing through CAB and
having the proper symmetries. Any other curve through these points passing
inside CAB will obviously not be convex. CAB is therefore a lower bound
for the yield loci.
Now draw a horizontal line symmetric about the a 3 axis intersecting the
adjacent axes of symmetry at C' and B'. Then it follows that no piecewise
smooth curve passing through C'AB' can lie outside C'AB', since then it
would not be convex. Thus C'AB' represents an upper bound to the yield
locus. The rest of the upper and lower bounds can be constructed from
symmetry and are shown as the two regular hexagons in Figure 6.3.7. Thus
It has been shown that all conceivable yield loci satisfying the conditions of
, equal yield in tension and compression, independence of hydro-
stress, and convexity must lie between two regular hexagons as shown;
convex curves are, of course, admissible. The usual Tresca yield locus is
represen1ted by the inner hexagon, and the von Mises circle circumscribes the
88 Criteria for Yielding [Ch. 6
inner hexagon and is circumscribed by the outer one. If the von Mises yield
surface is taken as a reference, then the maximum deviation of any admissible
yield surface is about 15.5 per cent.
To account for the influence of the intermediate stress in the von Mises
criterion, Lode introduced the parameter p.,, called Lode's stress parameter
(not to be confused with Poisson's ratio, p.,):
fk=
a2 - !(al + a 3)
!h- aa)
p, is thus the ratio of the difference between the intermediate stress and
average of the largest and smallest stresses to half the difference
the largest and smallest stresses, and is therefore a measure of the effect of
intermediate stress. Comparing (6.4.2) to (6.3.15) for the case a 1 :?:: a 2 :?::
we see that
p, = -v'3 tan()
Sec. 6-4] Lode's Stress Parameter. Verification of Yield Criteria 89
Thus if 8 varies from oo to 30°, f.L will vary from 0 to -I. All combinations
of loading from pure shear to simple tension are therefore included in the
range 0 to - I.
By means of (6.4.2), the von Mises criterion can be written
a 1 -as 2
(6.4.4)
ao = (3 + f.L2)1/2
The difference between the Tresca and von Mises criteria is then determined
by how much the right side of (6.4.4) differs from I. For f.L = -1, equation
(6.4.4) agrees with. (6.4.1). This is the case of simple tension, and the two
criteria agree as expected. The maximum difference in the range 0 to -1
obviously occurs when f.L = 0, which is the case of pure shear. The difference
is then 2/v'3, which agrees again with the previous results.
Lode ran a series of tests covering the range of f.L from 1 to -1 (- 30° :::; 0
:::; 30°). A f.L of 1 corresponds to uniaxial compression. The results are shown
in Figure 6.4.1. It is seen that the data favor the von Mises criterion even
1.30
I
• Steel
1.25 - o Copper
+ Nickel 0
1.20
0" 1-0"3 2
..J.----a:o- ="'3 +fJ \
n~1.15
/../ 0~
:f:\ "'~
le Oo
v J
1.10 u 0 0 :
·, ._,
1.00
1.00 v
-1.0 -0.8 -0.6 -0.4 -0.2
t o-;;;3 =\0\
0
fL
0.2 0.4 0.6
""
0.8
~ 1.0
+
cr1 lcro
0.4
Txy 03
cro .
0.2
• Copper
0.1 >< Aluminum
o Mild steel
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
crxlcr0
FIGURE 6.4.3 Results of Taylor and Quinney.
The theoretical curves are obtained as follows. Using the coordinate
of Figure 6.1.1, for tension and torsion loads only, the stresses are ax
Txy, all the others being zero. The Tresca criterion then becomes
2 J4a~ + 2
Txy = ao
or
Sec. 6-4] Lode's Stress Parameter. Verification of Yield Criteria 91
(6.4.6)
Both (6.4.5) and (6.4.6) plot as ellipses in the axrxy plane, as shown in Figure
6.4.3. It is seen in the figure that the data fit the von Mises criterion consider-
ably better than the Tresca criterion, although appreciable deviations from
the von Mises criterion do sometimes occur. Figure 6.4.4. shows a replot of
0 1.0
r---~--~---.----,----/~
/
/ 0
/
Tresca-.."'>/
/
/0
/
/
/
/0
/
/
/
/
/
/
/-
some of these results in the a 1 a 2 plane, and Figure 6.4.5 shows similar results
obtained by Ros and Eichinger [7]. Other tests of similar nature can be found
in the literature.
0 1.0
r---.----r--~------~
/
/
/
/o
/
/
/
/
/
/
h
Attempts have been made to improve the correlation of the data by includ~
ing the effect of the third invariant J 3 into the yield criterion [8]. It seems,
however, that from an engineering viewpoint the accuracy of the von Mises
criterion for yielding is amply sufficient, considering the general scatter and
lack of uniformity in the properties of nominally the same material obtained
from different batches from the manufacturer. Therefore, the search for more
accurate theories, particularly since they are bound to be more complex,
seems to be a rather thankless task. In what follows, therefore, we shall
generally use the von Mises yield criterion and occasionally the Tresca
criterion.
So far we have discussed the initial yield surface at which a material will
first start yielding. For a perfectly plastic material, this yield surface remains
fixed, as is seen in the uniaxial tensile test, where the stress after yielding re-
mains constant at the yield stress [Figure 2.6.1(d)). However, for a material
that strain hardens the yield surface must change for continued straining
beyond the initial yield. We know that for the uniaxial case, if a material is
strained beyond the yield point to some point such as B' in Figure 2.1.2, the
load removed so that the stress state moves to C', and then the load is in~
creased again, yielding will not take place until the point B' is reached again.
Thus the yield point has been raised in the work-hardened material. In
same way the yield surface in the case of multiaxial stress must "move or·
in some way, at least at the point where yielding initially took place.
In equation (6.3.1) we defined a yield function by the relation
such that whenever the function F became equal to the constant K, yielding
would begin. K then represented an initial yield surface in the stress space.
We can now generalize this type of relation to subsequent yield surfaces.
After yielding has occurred, K takes on a new value (or values), Cle}Jendmg
on the strain-hardening properties of the material. If the material is uu•.v"'""'"
and then loaded again, additional yielding will not occur until the new
of K is reached. The function F can then be looked upon as a loading ru11rn~nn;
or loading surface, which represents the load being applied, and the
K is a yield function, or strain-hardening function, and will depend on
complete previous stress and strain history of the material and its
Sec. 6-5] Subse'quent Yield Surfaces. Loading and Unloading 93
(1) F = K
oF
dF= ~da11 > 0
ualf
This constitutes loading.
oF
(2) F = K dF= -dalf
oajj
=0 (6.5.2)
=
oF
(3) F K dF= ~dalf < 0
ualf
existing state of stress lying on the yield surface. If the stress state changes to
Q, so that the vector da points" outward" from the cylinder, then loading is
taking place. If Q lies on the surface, we have neutral loading, and if Q is
inside the surface, we have unloading. As loading continues, the point
representing the new existing state of stress will again lie on the surface,
which will move and/or change its shape correspondingly. A similar picture
can be drawn in the 7T plane.
Assume the material obeys the von Mises criterion, so that the initial yield
locus is a circle of radius vi a 0 in the 7T plane. Suppose that straining takes
place to some point a~ > a 0 and the material is then unloaded. If we now
assume that the material remains isotropic, just as it was originally, then the
new yield locus is a circle of radius vi a~, which is larger than, but concentric
with, the original yield circle. We have thus assumed that the material strain
hardens isotropically, as shown for the uniaxial case by the curve ABCFG of
Figure 2.3.1. For isotropic hardening, therefore, the yield cylinder will expand
with stress and strain history but will retain the same shape as initially. For
a Tresca material, the subsequent yield loci will be a series of concentric
regular hexagons. This is shown in Figure 6.5.2.
·;;;
Q,
-10
-20
---
-25~~~--~~--~~--~~~
into account a Bauschinger effect but, because it maintains the total elastic
range constant, it probably overcorrects somewhat for the Bauschinger effect,
as discussed in Section 2.3. Attempts have therefore been made to improve
on the kinematic hardening model. This type of model, however, is much
more difficult to handle mathematically, and therefore the isotropic hardening
assumption is still generally used. For small plastic strains it probably gives
answers that are sufficiently accurate.
Problems
1. Compare the values of the pure shear yield strengths based on the six -·!!ories
of yielding listed (exclude Mohr's theory). Assume p, = 0.3 and a 0 ,c = 1;JSa0.
For this case, r = a1 = - a2.
2. A circular shaft 10 in. in diameter has a tensile yield strength of 90,000 psi
and a compressive yield strength of 120,000 psi. Determine the twisting
moment M 1 required to produce yielding based on:
(a) The maximum stress theory.
(b) The maximum shear theory.
(c) The distortion energy theory.
(d) The internal friction theory.
3. Derive equation (6.2.10).
4. It was shown in Section 6.3 that if the von Mises and Tresca criteria
assumed to agree for the case of uniaxial tension, then they will disagree
the case of pure shear and the von Mises yield circle will circumscribe
Tresca hexagon. Show that if it is assumed that the von Mises and
criteria agree for the case of pure shear, then they will disagree for the
of uniaxial tension and the Tresca hexagon will circumscribe the von
ellipse.
5. Derive the general form of the Tresca yield criteria as given by "'Y.''""'v'l
(6.3.8).
References 97
6. Derive the equation for the Tresca yield criterion in the sector between the
- a 1 and a 3 axes as was done in the text for the sector between the a 2 and
-a1 axes.
7. Derive equation (6.4.4).
8. Show that Lode's stress parameter /h = 1 corresponds to uniaxial compres-
sion, /h = -1 to uniaxial tension, and /h = 0 to pure shear.
9. Derive equations (6.4.5) and (6.4.6).
References
General References
Hill, R., The Mathematical Theory of Plasticity, Oxford Univ. Press, London,
1950.
Hoffman, 0., and G. Sachs, Introduction to the Theory of Plasticity for Engineers,
McGraw-Hill, New York, 1953.
Marin, J., The Mechanical Behavior of Engineering Materials, Prentice-Hall,
Englewood Cliffs, N.J., 1962.
CHAPTER 7
PLASTIC
STRESS-STRAIN
RELATIONS
In the previous sections the relations between stress and strain in the
elastic range were discussed and also the stress states at which plastic flow
or yielding will begin. There is one more ingredient necessary in constructing
a plasticity theory-the relations between stress and strain when plastic flow
is occurring. These relations are the subject of this chapter.
Whereas the strains are linearly related to the stresses by Hooke's law in
the elastic range, the relation will generally be nonlinear in the plastic range,
as is evident from the uniaxial stress-strain curve. A more complicated
distinction between elastic and plastic stress-strain relations arises from the
fact that whereas in the elastic range the strains are uniquely determined
by the stresses, i.e., for a given set of stresses we can compute the strains
directly using Hooke's law without any regard as to how this stress state
was attained, in the plastic range the strains are in general not uniquely
determined by the stresses but depend on the whole history of loading or
how the stress state was reached. This can readily be illustrated by considering
a thin-walled tube in tension and torsion as in the experiments of Taylor and
Quinney.
Consider the initial yield curve to be as shown in Figure 7.1.1. Let the
specimen be strained in uniaxial tension beyond the initial yield to some
98
Sec. 7-1] Distinction Between Elastic and Plastic Stress-Strain Relations 99
Txy
subsequent yield
E
8 A C a-x
point C, where CDE defines the subsequent yield curve. The plastic strains
will then be
e~ = ep
e~ = ef = -1-eP
e~y = e~z = efx =0
Let the specimen now be unloaded to the point B and let us apply a shear
stress increasing from B to D on the new yield locus. The plastic strains will
still be as given above. Any other path could have been used in arriving at
D from C such as OCFD, as long as we do not move outside the yield locus.
Now suppose that the specimen were first stressed in shear to the pointE on
the new yield locus and then, by any other path inside EDC, such as EGD,
were stressed to the point D. The plastic strains would be
e~y = ?'P
e~ = e~ = ef = e~z = e~z =0
which is obviously completely unrelated to the previous strain state. Thus
even though the same stress state at D exists for both loading paths, and
.!herefore the elastic strain states are the same, the plastic strain states are
different.
Because of the above illustrated dependence of the plastic strains on the
loading path, it becomes necessary, in general, to compute the differentials
or increments of plastic strain throughout the loading history and then
obtain the total strains by integration or summation. However, there is at
least one important class of loading paths for which the plastic strains are
independent of the loading path and depend only on the final state of stress.
These are the so-called radial or proportional loading paths, in which all the
--r
stresses increase in the same ratio. These will be more fully discussed subse-
quently.
I
7-2 PRA.NDTL-REUSS EQUATIONS
(7.2.1)
or
where S11 is the stress deviator tensor and d'A is a nonnegative constant which
may vary throughout the loading history. In these equations the total strain
increments are assumed to be equal to the plastic strain increments, the
elastic strains being ignored. Thus these equations can only be applied to
problems of large plastic flow and cannot be used in the elastoplastic range.
The generalization of equations (7.2.1) to include both elastic and plastic
components of strain is due to Prandtl [4] and Reuss [5] and are known as
the Prandtl-Reuss equations.
Reuss assumed that the plastic strain increment is at any instant of loading
proportional to the instantaneous stress deviation; i.e.,
(7.2.2)
or defJ = Su d'A
or de~= Sa d'A
or (7.2.4)
The numerators of the first three terms of (7.2.4) are the diameters of the
three Mohr's circles for the plastic strain increments and the denominators
are the diameters of Mohr's stress circles, as shown in Figure 7.2.1. ~9'~~t:vns
(7.2.4) therefore imply that the Mohr's circles of stress and plastic strain
jncre111en!_ are similar. Also from the relations for the principal shears,
equations (7.2.4) can be considered as stating that the ratios of the three
principal plastic shear strain increments to the principal shear stresses are
constant at any instant.
Equations (7.2.2) can be written in terms of the actual stresses as
FIGURE 7.2.1 Mohr's circles for stress and plastic strain increments.
dil =dyE
'~'oct
(7.2.8)
3
= v2 '~'oct
(7.2.9)
and
= v2 dyE
Sec. 7-2] Prandtl-Reuss Equations 103
For a uniaxial tensile test in the x direction the equivalent stress and equiva-
lent plastic strain increment reduce to
(7.2.11)
The convenience of the above definitions now becomes apparent. The equiva-
lent or effective stress, a., and the equivalent or effective plastic strain incre-
ment, dev, will henceforth be used in this text rather than the octahedral
shear stress and octahedral plastic shear strain increment.
The constant dl.. can therefore be written
(7.2.12)
(7.2.13)
p 3 dep
dByz = -2- Tyz
Ue
p 3 dev
dBzx = 2-- Tzx
Ue
or
p 3 dep S
dBjj = - - IJ
2 Ue
If one compares equation (7.2.9) for the equivalent stress u8 with equation
(6.2.8), which gives the von Mises yield criterion, it is seen that just as yielding
be~ins
(7.2.14)
where u 0 is the yield stress in simple tension. The equivalent stress is thus the
same as the von Mises yield function, and since equations (7.2.13) make use
104 Plastic Stress-Strain Relations [Ch. 7
of this function, the original Prandtl-Reuss assumptions imply the von Mises
yield criterion. This will also be shown subsequently from other considera-
tions.
It also follows from (7.2.14) that for a perfectlyplastic material the Prandtl-:-
R._~u~s eg_uaticmsmay bewritten
:d P
• eli= 23 dep
~ lj
S (7.2.15)
For a material that work hardens, however, a. may be greater than a 0 , and
it is now necessary to find the relation between the equivalent stress ae and
the equivalent plastic strain increment, dev· Before this is done, we introduce
the concepts of plastic work and the measures of work hardening.
dW = a11 dell
= (def, + defJ)
ali
But dW• = a11 def1 is recoverable elastic energy, whereas the plastic deforma-
tion is an irreversible process from which the energy cannot be recovered.
The remainder of the work done is called the plastic work per unit volume,
(7.3.2)
(7.3.3)
Now let us consider again a plot in the 7T plane where the axes are taken to
be the principal stress deviators S 1 , S 2 , and Sa (for a point lying in the
7T plane, since a 1 + a 2 + a a = 0, S 1 = a 1 , S 2 = a 2 , Sa = a a). Since for plastic
strain increments def + de~ + de~ = 0, the same plot in the 7T plane can
Sec. 7-3] Plastic Work. Two Measures of Work Hardening 105
be used for both the stress deviators and the plastic strain incrementsl pro-
vided the latter are multiplied by a constant to give them the dimensions of
stress. For this constant we choose 2G. The stress deviator vector and the
plastic strain increment vector can then be plotted on the same plot, as
Rdo
shown in Figure 7.3.1. From (7.3.3) the plastic work increment is the scaler
product of the two vectors, or
But
OP =.VSf + s~ + s~ =Viae
and
(7.3.4)
(7.3.5)
(7.3.6)
106 Plastic Stress-Strain Relations [Ch. 7
(7.3.7)
where
(7.3.9)
The functional relationship between the yield function and the plastic work
can be obtained experimentally and then the plastic strain increments can
be calculated using equation (7.3.6).
The second hypothesis uses the equivalent plastic strain previously defined
as a measure of work hardening; i.e., --
(7.3.10)
where dep is given by equations (7.2.10). The yield function is then assumed
to be a function of the equivalent plastic strain. Thus
Sec. 7-3] Plastic Work. Two Measures of Work Hardening 107
(7.3.12)
(7.3.13)
<re
e =E +ep
FIGURE 7.3.2 Relation between equivalent stress and equivalent plastic strain.
108 Plastic Stress-Strain Relations [Ch. 7
(7.4.1)
For the uniaxial tensile test in the 1 direction, aT becomes equal to a 1 • The
Prandtl-Reuss equations can now be written
(7.4.2)
with dev defined as previously. However, it can easily be shown (see Problem
5) that equations (7.4.2) are inconsistent with the definition (7.4.1). In spite
of this inconsistency, equations (7.4.2) are sometimes used to good advantage
[9]. Since in using the Tresca criterion it is assumed that the middle stress has
no effect on yielding, it is reasonable to assume also that there is no plastic
flow in that direction. A consistent set of relations can therefore be obtained
by assuming
d81p = v3dBp
T
ds~ =0
Equations (7.2.13) and (7.4.3) are known as the flow rules associated with
the von Mises and Tresca criteria, respectively. In Section 7.6 we will show
that these are special cases of a more general flow rule. In that section the
results of the previous sections will be essentially rederived based on a more
unified approach.
or v=p, (7.5.2)
-1 0
JL
FIGURE 7.5.1 Results of Lode's tests.
-1.0
FIGURE 7.5.2 Results of Taylor and Quinney's tests.
associated flow rule being used occasionally. It is to be noted that the assumed
coincidence of the principal stress axes and principal plastic strain increment
axes are borne out very well by the experiments of Taylor and Quinney.
or
dati( defi + defJ) > 0} (7.6.1)
dalf defJ:?: 0
or (7.6.3)
and from the linearity assumption 2 it follows that the superposition principle
may be applied to the stress and strain increments. Thus if dai1 and da71 are
two increments producing plastic strain increments, defJ' and de[j', then an
increment dalf = dai1 + da71 will produce an increment, de[j + de[j'.
Now assume that for a given state of stress a~c 1 , an increment of stress da~cz
producing plastic flow is imposed. This increment da~c 1 can be decomposed
into two parts da~ 1 and daZ 1 such that da~ 1 produces no plastic flow and daZ 1
is proportional to the gradient of f(ali). Geometrically this means that the
vector da~c 1 is decomposed into a component tangent to f and a component
perpendicular to j, as shown in Figure 7.6.1.
Since the increment dak 1 produces plastic flow, from (7.6.3) we have
-
8f
8 akl
J
uakl = -88f (-'~uaklI + uakl
J II)
> 0 (7.6.4)
akl
(7.6.5)
Also, da~ 1 has been taken proportional to the gradient off; therefore,
(7.6.6)
where a is a scaler > 0 and from (7.6.4), (7.6.5), and (7.6.6) it follows that
Hence
(7.6.7)
(7.6.8)
(7.6.9)
(7.6.10)
114 Plastic Stress-Strain Relations [Ch. 7
But da;1 produces no plastic flow, so that the increment dau = Cdaj1 + da;'1
for any value of C, positive or negative, will produce the same plastic incre-
ment defJ. We can therefore write the strain-hardening condition as
(7.6.11)
But daj1 defJ must vanish;, otherwise C could be chosen (a large negative
number) so as to violate (7.6.11). Therefore,
But
Hence
(7.6.12)
where G is a scalar which may depend on stress, strain, and history. Substi-
tuting (7.6.12) into (7.6.9) gives
(7.6.13)
or (7.6.14)
which is the general stress-strain relation consistent with the original assump-
tions.
Let us take some specific examples. For the von Mises yield function, let
!f = ~{al -
ua 1
!(a2 + as)]
Therefore,
Sec. 7-6] General Derivation of Plastic Stress-Strain Relations 115
f= ·Ha1- aa)
of= 0 of
oa2 oa3
Then
def =-! d'A
de~= 0
de~= --!-d'A
which are the same as equations (7.4.3). We note that the form of the flow
rule or plastic stress-strain relations associated with the Tresca criterion is
entirely different than that for the von Mises; thus each yield condition has
an associated flow rule, as was pointed out in Section 7.4. This is sometimes
ignored and, for example, the Tresca criterion has been used with the
von Mises flow rule. There is, however, no theoretical justification for that
type of assumption .
.\._\· . It is worth noting one ot.her importan.t fact from the previous derivation .
. Since da;, defJ = 0 and da;1 is tangent to the yield surface, it follows that
I "
l! defJ is normal to the yield surface, for the above equation merely represents a
h ..
dot product of two vectors. This can also be seen from (7.6.14), since defJ is
equal to the gradient of jtimes a scalar. It also follows from the above that !i
the Prandtl-Reuss equations imply the use of the von Mises criterion. .i
To summarize: Starting with the definition of work hardening and postu- '
lating the existence of a loading function and linearity between increments
of stress and increments of strain, we can arrive at the general flow rule
(7.6.14) for a strain-hardening material. It can also be shown that the plastic
strain increment vector must always be normal to the yield surface. The
scalar G, which depends in general on the stress, strain, and history, must be
determined from experiment, and its derivation will be discussed shortly.
For this case the work done by an external agency which slowly applies
and removes a set of stresses is zero over the cycle, or
(7.6.15)
116 Plastic Stress-Strain Relations [Ch. 7
It should be remarked that this equation is not the same as the second of
(7.6.1) with the equality sign. In (7.6.1) the equality sign is used only when
defJ = 0.
For ideal plasticity it is also assumed thatf(a11) exists and is a function of
stress only, and that plastic flow takes place without limit when f(a 11) = K
and the material behaves elastically when f(a 11) < K. For plastic flow,
therefore,
8/
df= -dali =0 (7.6.16)
8a 11
8
deft = dA / (7.6.17)
8a11
where dA. is a scalar.
then
Sec. 7-6] General Derivation of Plastic Stress-Strain Relations 117
n=2 c= 1/3
(7.6.19)
and since
(7.6.20)
(7.6.21)
and, iff= a 1 - a3 with a 1 > a 2 > a3 as for the Tresca criterion, then
(7.6.22)
where the incompressibility condition def + de~ + de~ = 0 has been used.
A second method for arriving at (7.6.21) is sort of intuitive. One seeks to
find a definition of effective plastic strain increment which when integrated
118 Plastic Stress-Strain Relations [Ch. 7
Therefore,
C= vt
dep = VidefJ defJ
and, for f = 12,
(7.6.24)
so that the integrated effective strain is a function of effective stress only; i.e.,
(7.6.25)
It should be noted here that the definition (7.6.21) for dev has been derived
for f = 12 only. Drucker has shown that it is reasonably correct for almost
any /(12 , 13). The second intuitive approach for defining dev is, of course, not
based on any specific loading function.
We are now in a position to determine the function G. It should first be
realized that for the previous formulation to agree with the uniaxial tensile
curve, da./dev must be the slope of that curve (in the plastic range). Substi-
tuting the basic equation
(7.6.26)
or (7.6.27)
or (7.6.29)
(7.6.30)
\Equations (7.6.30) constitute the flow rule (or plastic stress-strain relations)
1associated with the von Mises yield criterion. They are the well-known
Prandtl-Reuss relations we obtained previously. If we replace the plastic
strain increments in the above equations by total strain increments, the
Levy-Mises relations are obtained which are valid only if the plastic strains
are so large that the elastic strains can be neglected.
As a final note, a general flow law such as (7.6.14) can also be obtained on
the basis of a hypothesis that there exists a plastic potential (similar to the
strain energy density function) which is a scalar function of stress, g(a 11), from
which the plastic strain increments can be obtained by partial differentiation
with respect to the stresses. Thus
where df3 is a nonnegative constant. The plastic potential g(ali) was first
introduced by Melan [14]. By comparison with (7.6.14), it would appear that
the plastic potential should play the same role as the yield function, and
indeed Bland [15] has proved that they must be the same function, so that gin
(7.6.31) can be replaced by f; (7.6.31) and (7.6.14) are then the same.
whereby the total strain components are related to the current stress. Thus,
instead of (7.6.30), one would have
(7.7.1)
The plastic strains then are functions of the current state of stress and are
independent of the history of loading. Such theories are called total or
clf![ormation theories in contrast to the incremental or flow theories previously
described. This type of assumption greatly simplifies the problem; however,
as was previously shown, the plastic strains cannot in general be independent
of the loading path and deformation theories cannot generally be correct.
There has often been a tendency therefore to ignore all deformation theory
as of little value.
It can easily be shown, however, that for the case of proportional or radial
loading, i.e., if all the stresses are increasing in ratio, the incremental theory
reduces to the deformation theory. For if a11 = Ka~, where aZ· is an arbitrary
reference state of stress (nonzero) and K is a monotonically increasing func-
tion of time, then S11 = KS?1 and a. = Ka~ and (7.6.30) becomes
3dep
dBt·P1 = - so
- i}
2a~
(7.7.2)
so the plastic strain is a function only of the current state of stress and is
independent of the loading path.
Furthermore, it has been proposed by Budiansky [17] that there are ranges
of loading paths other than proportional loading for which the basic postu-
lates of plasticity theory are satisfied by deformation theories. Budiansky's
theory postulates the occurrence of corners or singular points on the succes-
sive yield surfaces and, although the existence of such singular points has as
yet not been established experimentally, one cannot rule out the possibility
of loading paths other than proportional loading for which total plasticity
theories may give satisfactory answers.
From a practical viewpoint, there are a great many engineering problems
where the loading path is not far from proportional loading, provided one is
careful when unloading occurs to separate the problem into separate parts, the
loading parts, and the unloading parts.
Sec. 7-8] Convexity of Yield Surface. Singular Points 121
Problems of plastic flow in thermally stressed disks and cylinders have been
handled in this way and good results obtained using deformation theory.
On the other hand, it will subsequently be shown that with the present
widespread availability of high-speed computors, many simplifying assump-
tions heretofore made, including the use of deformation theories under
doubtful conditions, are often unnecessary.
In Section 6.3 the statement was made that the yield surface was convex.
A proof, as given in reference [13], will now be presented. Consider some state
of stress a~ inside the loading surface, as shown in Figure 7.8.1. Let some
external agency add stresses along some arbitrary path inside the surface
until a state of stress au is reached which is on the yield surface. Only elastic
changes have taken place so far. Now suppose the external agency to add a
very small outward pointing stress increment dati which produces small
plastic strain increments defJ, as well as elastic increments. The external
agency then releases the dati and the state of stress is returned to a~ along an
elastic path. The work done by the external agency over the cycle is
(7.8.1)
a11 - at and the vector defJ plus the scalar product of dali and de{j. Now,
from the strain-hardening definition equation (7.6.1),
That is, the vectors dali and deft make an acute angle with each other. In a
similar fashion, since the magnitude of a11 - at can always be made larger
than the magnitude of da11 , it follows that
(ali - a~)defJ ~ 0
or laii - a~lldefJI COS rp ~ 0
Hence
1T 1T
-- < .1. <-
2-'f'-2 (7.8.3)
Thus the vector ali - at makes an acute angle with the vector deft for all
choices of at. Therefore, all points a~ must lie on one side of a plane per-
pendicular to deft, and, since deft is normal to the yield surface, this plane will
be tangent to the yield surface. This must be true for all points ali on the yield
surface, so that no vector a11 - at can pass outside the surface intersecting
the surface twice, as shown in Figure 7.8.3. The surface must therefore be.
convex. On the other hand, if the surface is not convex, there exist some
a
po1n:ts 11 and at a at
such that the vector 11 - forms an obtuse angle with
the vector da1J> as shown in Figure 7.8.4. This completes the convexity proof.
Equation (7.6.14) implies that the yield surface has a unique gradient. It
may happen, however, that the yield surface has vertices or corners where the
gradient is not defined. For example, the Tresca hexagon has no unique
normal at the corners, where two of the stresses are equal. Such points are
called singular points or singular yield conditions. Such points can be treated
by introducing an auxiliary parameter as described in reference [18].
(7.9.1)
where ef1 is the elastic component of the total strain, ef, is the accumulated
plastic strain up to (but not including) the current increment of load, and
..........------------------
124
Plastic Stress-Strain Relations {Ch. 7
1
~sfJ is the increment of plastic strain due to the increment of load. sfJ is
presumed to be known, ~sfJ is to be computed. Define modified total strains
as follows:
(7.9.2)
Then
(7.9.3)
Subtracting the mean strain from the diagonal components of both sides of
equation (7.9.3) results in
(7.9.4)
where er, is the elastic strain deviator tensor and e;, is the modified strain
deviator tensor. From Hooke's law and the Prandtl-Reuss relations,
1 S 1
e,,e = 2G 11 = 2G ~A us,,
A P
Hence
(7.9.5)
(7.9.6)
A p D.Bp (
LJ.Bx = 3- 2BxI -
I
By -
I)
Bz
Bet
A p D.Bp (
LJ.By = - 2ByI -
I
Bz -
1)
Bx
3Bet
A p D.Bp (
U.Bz = 3- 2BzI -
I
Bx -
I)
By
Bet
(7.9.10)
Bet
V2 [(BxI
= T -
1)2
By + ( I
By -
1)2
Bz + ( I
Bz -
I
Bx
)2 + 6( I
Bxy
)2 + 6(I
Byz
)2 + 6( I
Bzx
)2]1/2
(7.9.11)
and the primed quantities are the modified total strains as given by equation
(7.9.2).
Equations (7 .9 .1 0) are equivalent to the Prandtl-Reuss equations. The
stresses do not appear in these equations and the increments of plastic strain
can be computed from the total strains. Note that since they have been
derived by use of the Prandtl-Reuss equations, they implicitly make use
of the von Mises yield function. It should also be emphasized that the equiva-
lent total strain defined by (7.9.7) is a purely mathematically defined quantity
without any direct physical meaning, even in the uniaxial case. However, it
can be related to the uniaxial stress-strain curve as follows: From equation
(7.2.12),
(7.9.12)
126 Plastic Stress-Strain Relations [Ch. 7
1 O'e ) Set
( + 3G L\sp = L\sp
or
(7.9.13)
(7.9.14)
ae = Ue,i-l + (ddae).
Sp 1-1
L\ep + ...
Sec. 7-10] Complete Stress-Strain Relations. Summary 127
where
Bet = V je11eu
(7.9.17)
for each of the load increments. All these increments of plastic strain are then
I
added to give the total plastic strain. The integration is thus replaced by a
summation.
Let the total loading path be divided into N increments of load. Assume
that the plastic strains have been computed for the first i - 1 increments of
load and we now wish to compute them for the ith increment of load. The
total strains at the end of the ith increment can be written with thermal
strains included, as
(7.10.1)
(7.10.2)
In the above equations the sums are known and the problem is to calculate
the plastic strain increments for the current or ith increment of load, and the
corresponding stresses. To do this it is necessary to use one or another of the
plastic stress-strain relations discussed in previous sections. A yield criterion
must be chosen and the associated flow rule as given by equation (7.6.14).
In particular, since we shall concern ourselves only with the von Mises and
Sec. 7-10] Complete Stress-Strain Relations. Summary 129
Tresca yield criteria, the following relations previously derived will be used.
For the von Mises criterion:
ae = J
v'3:J; = 2 Tact
= V!St}Stj
(7.10.3)
(7.10.4)
or
(7.10.5)
Alternatively we define
Plastic Stress-Strain Relations [Ch. 7 l
1-1
e;, = el, - 2:
k=1
ilB&,k = B;, - s,,Bm
1-1
BtJ = B11 - L
k=l
ilBfJ,k
= v2 [(
3 I
Bx - ByI )2 + ( ByI - Bz1)2 + ( BzI - I )2
Bx
(7.10.6)
Then
or
A p
UBx = -!1Bp
3
(
2 BxI -
I
By -
I)
Bz
Bet
A p
UBy = -
!1Bp (2 ByI -
I
Bx -
I)
Bz .
3Bet
A P
UBz = -!1Bp
3
(
2BzI -
I
Bx -
I)
By
Bet
(7.10.7)
where a 8 , 1 _ 1 is the value of the equivalent stress at the end of the (i - 1)st
increment of load and (da./dep) 1 _ 1 is the slope of the uniaxial tensile curve
replotted as true stress versus true plastic strain. _Equation (7.10.9) is exact
for)inear_strain hardening. The above relations are shown graphically in
Figure 7.9.1.
If the deformation or total theory of plasticity is used, all the above
relations are valid if the Ll's are removed from all the previous equations
and the primes are removed from equations (7.10.6) through (7.10.7).
Equation (7.10.8) becomes
(7.10.10)
and by the use of (7.10.10), the uniaxial stress-strain curve can be replotted
as a curve of eP versus e 81 as shown in Figure 7.10.1. This curve can then be
used instead of the original stress-strain curve.
24 ~10-3 /
....
~ 20
c
·et; 16
-+-
c
/
~ 12
>
·:; /
cr /
Q) 8
'E /
~
4
v
0 4 8 12 16 20 24X10- 3
Equivalent plastic strain, ep
(7.10.11)
or, alternatively,
and the relation between aT and Llep is taken from the uniaxial tensile curve.
132 Plastic Stress-Strain Relations [Ch. 7
The stress-strain relations discussed in this chapter are just one of four
sets of relations that must generally be satisfied in solving an elastoplastic
problem. The other three sets of relations are the same as for any elasticity
problem. These are
1. The equations of equilibrium of stresses.
2. The strain-displacement or compatibility relations.
3. The boundary conditions.
To obtain a complete solution we must find a set of stresses and strains
which satisfy these four sets of relations. In the next several chapters it will
be shown how these relations are adapted to specific problems and how
solutions to these problems can be obtained. In all that follows, as in the
preceding, it is assumed that the material is homogeneous, isotropic, and
strain hardens isotropically.
Problems
1. Show that the equivalent stress a 6 defined by equation (7.2.9) can also be
written
a.= VJStJSij = VJ(Si + s~ + S5
2. Show that the equivalent plastic strain increment dev defined by equation
(7.2.10) can also be written
dep = vt [(de~) 2
+ (de~) 2 + (de~) 2 + 2(de~y) 2 + 2(de~.) 2 + 2(de~x)2] 1 1 2
= ~'3 [(de~) 2 + (de~) 2 + de~ de~ + (de~y) 2 + (de;2 ) 2 + (de~x) 2] 1 1 2
v2
= - - [(de~ - de~) 2 + (de~ - den 2 + (de~ - deD 2
3
+ 6(de~y) + 6(de;2 ) 2 + 6(de~x)2]112
2
8. From the fact that the plastic strain increment vector is normal to the yield
surface, prove that the Prandtl-Reuss equations imply the use of the
von Mises yield criterion.
9. Derive equation (7.6.30) from equation (7.6.28).
References 133
10. Show that the Prandtl-Reuss relations imply that the principal axes of
stress and of plastic strain increment coincide.
11. Derive equations (7.2.12) and (7.2.13) using tensor notation only.
12. Determine the equivalent stress aT for the Tresca criterion by means of
equation (7.6.18). Assume a1 > a2 > as. Also determine the effective plastic
strain increment by the two methods described by equation (7.6.20) and
what follows.
13. Prove that
References
15. D. R. Bland, The Associated Flow Rule of Plasticity, J. Mech. Phys. Solids,
6, 1957, pp. 71-78.
16. H. Z. Hencky, Zur Theorie Plastischer Deformationen und der hierdurch
im Material hervorgerufenen Nachspannungen, Z. Angew. Math. Mech.,
4, 1924,pp. 323-334.
17. B. Budiansky, A Reassessment of Deformation Theories of Plasticity,
J. Appl. Mech., 26, 1959, pp. 259-264.
18. W. T. Koiter, Stress-Strain Relations, Uniqueness, and Variational Theo-
rems for Elastic-Plastic Materials with a Singular Yield Surface, Quart.
Appl. Math., 11, 1953, pp. 350-354.
19. A. Mendelson and S. S. Manson, Practical Solution of Plastic Deformation
Problems in Elastic-Plastic Range, NASA Tech. Rept. R-28, 1959.
General References
Drucker, D. C., Stress-Strain Relations in the Plastic Range, a Survey of Theory
and Experiment, Office of Naval Research, Contract N7 onr-358, NR-()41-()32,
Dec. 1950.
Hill, R., The Mathematical Theory of Plasticity, Oxford Univ. Press, London,
1950.
Johnson, W., and P. B. Mellor, Plasticity for Mechanical Engineers, Van Nostrand,
Princeton, N.J., 1962.
Naghdi, P. M., Stress-Strain Relations in Plasticity and Thermoplasticity, Office
of Naval Research, Contract Nonr-222 (69), Tech. Rept. No. 9, 1960.
CHAPTER 8
ELASTOPLASTIC
PROBLEMS OF
SPHERES AND
CYLINDERS
Spheres and cylinders are widely used as pressure vessels, in the chemical
industry, for example, as well as many other places. The loads involve high
pressures and sometimes high temperatures and high temperature gradients.
The elastic stress and strain distributions are relatively simple to obtain,
particularly since the loading is usually reasonably symmetric. The solutions
in the elastoplastic range, however, become complicated, and so simplifying
assumptions of various types are made. These usually involve assuming the
material to be incompressible in both the elastic and plastic ranges, and
assuming it to be perfectly plastic in the plastic range. With these assumptions
closed-form solutions can be obtained. We shall first present some of these
classical solutions. Subsequently it will be shown how these problems can be
solved without the usual simplifying assumptions.
For later use we record here the equilibrium, compatibility, strain-
displacement, and stress-strain relations in spherical coordinates and polar
coordinates assuming spherical and axial symmetry, respectively.
Spherical Coordinates
The stresses are designated by a, and a 8 = a,p and the strains by e, and
e8 = e,p. The equilibrium equations reduce to
(8.1.1)
135
136 Elastoplastic Problems of Spheres and Cylinders [Ch. 8
where Fr is the body force per unit volume. The strains are related to the
displacements by
du u
e =-
r dr Be= f = Bq, (8.1.2)
Because of symmetry the shear stresses and shear strains are zero as well as
the tangential displacements.
The stress-strain relations are
1
er = E (ar - 2p..ae) + aT+ e{.'
(8.1.4)
1
Be = E [(1 - p..)ae - p..ar] + aT + e~
where e;: and e~ are the total plastic strains. From the incompressibility
condition it follows that
e{.' = -2e~ (8.1.5)
For the von Mises yield criterion, the equivalent stress becomes
(8.1.6)
so that the yield criterion is
(8.1.7)
derp = 2d d ar - ae
- eo
P
= eP Iar- ae I
= dep sgn (ar - ae)
Sec. 8-1] General Relations 137
where sgn stands for "the sign of." We note that if the plastic strains vary
monotonically with the applied load, equation (8.1.9) can be integrated to
give
(8.1.10)
Note also that the Tresca yield criterion in this case coincides with the
von Mises criterion.
Polar Coordinates
We assume axial symmetry and either plane strain or plane stress. The
equilibrium equations then become
(8.1.11)
where Fr is the body force per unit volume, The strain-displacement relations
and corresponding compatibility equation are
(8.1.12)
which are the same as equations (8.1.2) and (8.1.3). The stress-strain relations
are given by
(8.1.13)
For the case of plane stress, uz = 0, and for the case of plane strain ez = 0
or ez = constant for generalized plane strain. In both cases the shear stresses
and strains are zero.
The von Mises and Tresca criteria do not coincide in this case as they do
for the case of spherical symmetry. The yield criteria and corresponding
plasticity relations will be described subsequently as they are used. Several
examples will now be discussed beginning with the case of a thick hollow
sphere.
138 Elastoplastic Problems of Spheres and Cylinders [Ch. 8
ar = --2E 1
1 --1-L-
r3
J' aT r dr + -23 ( 1 -
a
2
3
-ar3 ) C1 + -Cr32
(8.2.1)
where C1 and C2 are integration constants. Note that E and a have been
assumed constant in obtaining the above solution. The constants C1 and C2
can be obtained using the boundary conditions
ar(a) = -p
(8.2.2)
ar(b) = 0
resulting in
c2 = -pa 3
_ 3E
Cl - -1-- b3
1
3
fb aT r2 dr + -23 b3pa3 3 (8.2.3)
-1-L -a a -a
For convenience the following dimensionless quantities are now introduced:
b r P=p
{3=-a p=-a
ao
(8.2.4)
EaT = ar
S r- ae
T = (1 Se =-
- /-L)ao ao ao
where a 0 is the yield stress in uniaxial tension. Equation (8.2.1) can now be
written in the dimensionless form as
(8.2.5)
Sec. 8-2] Hollow Sphere with Internal Pressure and Thermal Loading 139
The strains can be computed from (8.1.4) and the displacement from
(8.1.2).
We note that in the case of pressure loads only, the stress distribution is
independent of Poisson's ratio [see equations (8.2.5) and (8.2.4)]. The as-
sumption that is often made that the material is incompressible in the elastic
range (p, = -!), as well as the plastic range, therefore leads to no error in the
elastic stress distribution. In the case of temperature loads, however, assuming
p, = t instead of 0.3, for example, results in approximately a 30 per cent
error in the elastic thermal stresses. The strains are not independent of
Poisson's ratio even for the case of pressure loading. In what follows, the
effect of Poisson's ratio is always taken into account.
The conditions for the onset of yielding in the sphere can now be investi-
gated. In terms of the dimensionless stresses defined in (8.2.4), the yield
criterion (8.1.7) is written
T = T 0a (~ _ 1)
b- a r
or 'T =~ (~-
f3- 1 p
1) (8.2.7)
EaT0
where To = (1 - p,)ao
Evaluating the integrals and substituting into the yield condition (8.2.6)
results in
Consider first the case of pressure only. Then the yield condition becomes
3[3 3P
2p3(f33 - 1) = 1 (8.2.9)
140 Elastoplastic Problems of Spheres and Cylinders [Ch. 8
Yielding will first occur at the smallest value of p, i.e., p = 1, and the dimen-
sionless pressure necessary to first cause yielding, the critical pressure, will be
2(fi3- 1)
Porlt = 3[33 (8.2.10)
A plot of the ratio of this critical pressure as a function of the ratio of the
outer to the inner radii f3 is shown in Figure 8.2.1. For a given value of {3,
yielding will start at the inner surface at a pressure as given by equation
(8.2.10) or Figure 8.2.1. AsP is increased, the plastic zone will spread from
Peri!
0.8
2/3
0.6 t:-------===:::::=====
0.4
0.2
the inner surface toward the outer surface. Note that as f3 approaches infinity
Porit approaches t, so that if the pressure is equal tot the yield stress, yielding
is sure to take place, no matter what the dimensions of the sphere.
Considering the case of temperature only, equation (8.2.8) gives
(8.2.11)
For this case yielding will also first start at the inner surface. However, if
both pressure and temperature are present, yielding may start at any radius,
depending on the relative values of P, T 0 , and {3.
As an example let f3 = 2. Then for the case of internal pressure alone,
equation (8.2.10) gives
7
Porlt = 12 = 0.583
To,orlt = 1.4
Sec. 8-3) Hollow Sphere. Spread of Plastic Zone. Pressure Loading Only 141
Porlt = 0.75
compared to 0.583 for pressure alone. The effect of the temperature gradient
in this case has been to retard the onset of yielding.
A complete discussion of the effects of temperature and pressure on
yielding is given in reference [1].
So far, only the start of yielding has been considered. The spread of the
plastic zone through the sphere is investigated next. The pressure problem
and the temperature problem will be discussed separately in Sections 8.3, 8.4,
and 8.5 under the assumption that the material is perfectly plastic. The general
solution for strain-hardening materials under combined pressure and thermal
gradient is presented in Section 8.6.
When only internal pressure is acting, yielding will begin at the inner
surface at a pressure given by equation (8.2.10); i.e.,
(8.3.1)
As the pressure increases, the plastic zone will spread outward toward
the outer surface. Let the radius to the end of the plastic zone be r 0 • Since the
material is assumed perfectly plastic, at every point in the plastic region the
equivalent stress is equal to the yield stress and since for this case a8 >a.,
(8.3.2)
142 Elastoplastic Problems of Spheres and Cylinders [Ch. 8
in the plastic region. Substituting into the equilibrium equation (8.1.1) gives
dSr 2
dp =p
or Sr = 2ln p +C
But at
p= 1 Sr = -P
Therefore,
C = -P
and S, = 2ln p- P }
P < Pc (8.3.3)
8 8 = (21n p + 1) - P -
Equations (8.3.3) give the stresses in the plastic region. Note that no stress-
strain relation was needed to obtain these stresses. The problem is therefore
called statically determinate.
At the plastic zone boundary, i.e., at p = Pc, the radial stress is
We can now consider the elastic part of the sphere as a new sphere with
inner radius r c and outer radius b, with an internal pressure given by equation
(8.3.4). Since at this new inner radius the sphere is just at the yield point,
equation (8.3.1) must apply with {3 replaced by f3c = b/rc, and -Pcrit replaced
by Sr,c· Hence
2 f3c 3 - 1
2lnpc- P = -3~
or P = 2 ln Pc + i( 1- ;~)
= 2ln ~
a
+ ~3 (1 - br~)
3 (8.3.5)
Equation (8.3.5) gives the pressure required to cause the plastic zone to reach
a radius rc or, alternatively, for a given internal pressure P, equation (8.3.5)
could be solved for the plastic zone radius rc. A plot of the pressure versus
the plastic zone radius is given in Figure 8.3.1 for f3 = 2.
When rc becomes equal to b, the sphere is completely plastic. This will
occur at a pressure [from (8.3.5)]
P = 2ln {3 (8.3.6)
Sec. 8-3] Hollow Sphere. Spread of Plastic Zone. Pressure Loading Only 14 3
Pc
FIGURE 8.3.1 Plastic zone radius versus applied pressure, f3 = 2.
S, = 2 ln p - 2 ln f3
= 2ln~
(8.3.7)
r
= 2ln b
r
S 0 = 1 + 2ln b
or
(8.3.8)
144 Elastoplastic Problems of Spheres and Cylinders [Ch. 8
Equations (8.3.3) and (8.3.8) give the complete stress distribution in the
sphere for given ratio of applied internal pressure to yield strength, with the
plastic zone boundary Pc given by equation (8.3.5).
To calculate the strains and displacements in the sphere, the stress-strain
relations and strain-displacement relations are used. For convenience we
define, as was done for stresses in equations (8.2.4), "dimensionless" strains
and displacements as follows:
(8.3.9)
To compute the strains before yielding begins, equations (8.2.5) (with tem-
perature terms deleted) are substituted into the stress-strain relations (8.1.4),
resulting in
(8.3.10)
(8.3.11)
When yielding starts P is equal Pcrit given by equation (8.2.10), and the
displacement at the inner surface, p = 1, is
2 {3~ - 1
Pcr!t =3 -r
Sec. 8-4] Hollow Sphere. Residual Stresses. Pressure Loading 145
acting at the inner radius rc. From (8.3.11) the displacement at r:::: rc is
obtained by replacing f3 by f3c and p by rfrc, resulting in
3
2 [ (1 - 2/k)P
U = 3{3~ + -l+ftf3
2 - p2 ] (8.3.13)
When the plastic zone reaches the outer radius b, p = [3, f3c = I, and
(8.3.14)
If the pressure is removed from the sphere discussed in Section 8.3 after
plastic flow has occurred over part of the sphere, residual stresses will result.
To find the residual stresses it is necessary to superpose on the stress system
due to the internal pressure p and temperature T a completely elastic stress
system due to a pressure -p and temperature -T. This will be correct as
long as yielding in reverse does not occur; i.e., the residual stresses are not
large enough to produce yielding. To see this, consider two stress systems
satisfying the following two sets of equations:
I I I
dar 2 ar - ae- 0
-+
dr r
-
de 8 e~ - e~
dr = - r -
(8.4.1)
a;(a) = -p
a;(b) = 0
146 Elastoplastic Problems of Spheres and Cylinders [Ch. 8
and
II II II
dar 2 ar - ao- 0
-+
dr r -
ll II II
dBr Br - eo
dr = --r-
(8.4.2)
eoII = E
1 [(1 - P,) aoII - p,arII] - a T
a;(a) = p
a;{b) = 0
The primed system corresponds to the system of stresses in the sphere with
temperature T and internal pressure P. The double-primed system corre-
sponds to the stresses in a sphere with temperature - T and internal pressure
-P. If the two systems are added together, there is obtained a system of
stresses ar = a~ + a;, etc., satisfying the following equations:
(8.4.3)
Thus the resultant system corresponds to the unloaded sphere with permanent
plastic strains due to the first system. If plastic flow occurs during the un-
loading, the elastic double-primed system can no longer be added to the
original system, but it is necessary to solve another plastic flow problem for
the new plastic strains.
For the case of pressure loading only, the elastic stresses due to a pressure
equal to - P are, from (8.2.5),
2p3 + [33
Sll- P
8 - - 2p3(!33 - 1)
Sec. 8-4] Hollow Sphere. Residual Stresses. Pressure Loading 147
Adding to the stresses given by (8.3.3) and by (8.3.8) gives for the residual
stresses
p3 - (33 )
s~ = 2ln p - p ( 1 + p3((33 - 1)
= 2 In p - ~3 Pcrlt
_!_ ( 1 - \)
P
or s; = ~3 [3ln p - _!_
Porit (1 - ~)]
P
}
p < Pc (8.4.5)
s~ = ~3 [~2 + 1
3ln p - _!_ (1 + 2 p3)]
Pcrit
-
s; = j [(P: 1
t - P~) {: 3 - ; 3 )] }
2 P 1 1 p :::0: Pc
(8.4.6)
S~ = -3 [(Pcrit - P~ )(2p + (3 )]
3 3
(8.4. 7)
an~ since P :::0: Parit• a residual compressive stress results. Upon reapplication
of a pressure less than or equal to the original maximum, only elastic strains
will occur. The shell has thus been strengthened by the initial pressurization.
If the material work hardens, an even greater strengthening can be achieved.
In the above derivation it has been assumed that no plastic flow takes place
during the unloading; i.e., there is no yielding in compression due to the
residual stresses. If such yielding occurs, then not only is our assumption
that the unloading is elastic violated, but the situation may be dangerous
with regard to the safety of the sphere. The maximum value of applied
pressure P such that if the sphere is unloaded there will be no reversed plastic
flow is called the shakedown pressure, P•. This pressure can be found as
follows. For reversed yielding the yield criterion can be written
s;- s~ = 1 (8.4.8)
Ps
1.6
4/3t------=:::::::::=====
1.2
FIGURE 8.4.1 Variation of shakedown pressure with thickness ratio for hollow
sphere with internal pressure.
As long as the applied pressure is less than twice the critical pressure, the
residual stresses will be elastic. Making use of equation (8.3.1) the shakedown
pressure can be written directly as a function of the thickness ratio f3:
(8.4.10)
(8.5.1)
2.0
1.6
+-
'Ei. 1.2
~ ~----------------
0.8
0.4
234567 8
{3
2({33 - 1) 2((3 2 + (3 + 1)
= (3(2(32 _ (3 _ 1) (8.5.2)
To,orit
(3(2(3 + 1)
dS, 2
dp-- p
or S, = -2ln p +C
and since S,(1) = 0, C = 0. Hence
S, = -2lnp }
(8.5.4)
S0 = -1-2ln p P <
-
Pc
Note that the stresses in the plastic region are independent of the temperature.
The radius of the plastic zone, r 0 , of course, depends on the temperature.
dEe + Ee - Er = 0
dp p (8.6.1)
Er = Sr - 2/hSe + (1 - fh)T + Ef
Ee = (1 - !h)S8 - fhSr + (1 - f.L)r + E~
Substituting the last two of equations (8.6.1) into the second, combining
with the first, and integrating results in the following equations:
(8.6.2)
Sec. 8-6] Hollow Sphere of Strain-Hardening Material 151
(8.6.3)
(8.6.4)
where
(8.6.5)
Sr(I) = -P
(8.6.6)
S//3) = 0
For the elastic case equations (8.6.2) reduce to (8.2.5). For the case of a
perfectly plastic material, the solution was given in the previous sections. We
shall consider here only the case of a strain-hardening material.
To obtain a complete solution to the problem, it is necessary to determine
the plastic strain distribution E~ through the sphere. This will, of course,
depend on the stress-strain curve of the material. The plastic strain distri-
bution can be obtained in the following manner. The equivalent stress is
related to the equivalent plastic strain through the stress-strain curve of the
material. Thus
lSI ;::: I
(8.6.7)
)SI ~I
(8.6.8)
2.0
p
c 1.5
p
FIGURE 8.6.1 Variation of plastic zone radius with applied pressure for different
temperature gradients:
f3 = 2, T = -r 0 (f3/p - 1)/(/3 - 1), ae = 30,000 + 136,000 ev 1' 2 •
Sec. 8-6] Hollow Sphere of Strain-Hardening Material 153
FIGURE 8.6.2 Variation of plastic strain with radius for various pressures:
{3 = 2, To = 0.
was used to perform the integrations. The cases shown are for illustrative
purposes only. Any combination cf geometry, loading, and material proper-
ties can be used and a rapid solution obtained. The time required to obtain
a complete solution for a given loading condition, using a high-speed digital
computer, is on the order of a few seconds.
This type of successive approximation method will be discussed at greater
length in Chapter 9, where several numerical examples will be given. Right
now it will be shown that if the material strain hardens linearly the solution
can for some cases be obtained in closed form.
For linear strain hardening it follows from Figure 8.6.3 that
Ev = 1- m (I Sl - 1) (8.6.9)
m
1-m S
= --(1-ISj)- 1-m(S - s )
Ep
r m lSI = -m- -
lSI (8.6.10)
Consider the case of pressure loading only. a8 will always be positive and
ar will always be negative, so that a 8 - ar > 0. Therefore, S/1 S I +1 and
1-m
Ef = ---m (1 - S) (8.6.11)
154 Elastoplastic Problems of Spheres and Cylinders [Ch. 8
"e
~ e
FIGURE 8.6.3 Stress-strain curve for linear strain hardening.
If the plastic zone extends to p = Pc• then, from (8.6.1I), making use of the
first of equations (8.6.1), it follows that
I- m I- m
Pr;P
--lnp- --(S
m 2m r
+ P) P ~ Pc
...!.. dp = (8.6.I2)
fl p { I - m In
m
_ I - m (S + P) P;::: Pc
Pc 2m r,c
where Sr,c is the value of S, at p = Pc· Substituting into (8.6.3) and (8.6.4)
results in
(8.6.I3)
I- m 3
S = 2(1 - tL)m (1 - S) + 2ps Cl (8.6.14)
(8.6.15)
(8.6.16)
which is the same result previously obtained for the perfectly plastic material.
Obviously the onset of yield depends only on the yield stress. AsP is increased,
the plastic zone spreads to Pc and -Sr,c can be considered to be the critical
pressure acting on a sphere with inner radius Pc and outer radius fie = fJ/ Pc·
Thus
(8.6.18)
Hence
2 fJ3 · [( 1 I - m ) 1- m 1 - m fJ~ - 1]
- p,)m P - (1 - p,)m In Pc - 3(1 - p,)m ~
3
3 Pc = fJ 3 - 1 + 2(1
(8.6.19)
or
p _ 4(1 - p,)m[(fJ3 - 1)/{13 ] p~ + 2(1 - m) ln Pc + i(l - m)(fJ~ - 1)/fJ~
- 2m(l - p,)+ (1 - m)
(8.6.20)
which relates the pressure P to the plastic zone radius Pc· Note that if m = 0,
2{1~-1
P = 2lnp 0 + 3~
which is the value previously obtained for the perfectly plastic material
[equation (8.3.5)].
As an example, for fJ = 2, m = 0.1, and p, = 0.3, the pressure required
for yielding of the complete sphere, Pc = fJ, is 1.83, compared to 1.39 for a
perfectly plastic material. It thus takes a 32 per cent higher pressure for the
strain-hardening sphere to yield completely as compared to the perfectly
plastic sphere.
To obtain the stresses we substitute into equations (8.6.2). Thus
Note that if m = 0 these reduce to the previously obtained values for the
perfectly plastic material. Thus to obtain the complete stress distribution,
Pc (or P) is obtained from (8.6.20), Sr,c from (8.6.18), and then the stresses
from (8.6.21) and (8.6.22).
EaT EaT0 r b
p=- (3=-
T = (1 - tL)ao y = (1 + tL)ao a a
where a and b are the internal and external radii, a 0 is the yield stress, and e0
is the yield strain.
Sec. 8-7] Plastic Flow in Thick-Walled Tubes 157
dSr Se- Sr
dp = - p -
(8.7.2)
If the Tresca criterion and its associated flow rule are used, then, assuming
Se > Sz > S,
(8.7.4)
and S = Se- Sr = 1 at yielding
Sr(a) = -P
(8.7.5)
Sr(b) =0
and the conditions at the end of the tube are determined by case I, 2, or 3
above. For plane strain Ez = 0. For generalized plane strain, Ez is a constant
which can be determined from the end loads on the tube. Thus let the axial
force acting on the tube be F. Define
F* = J: Szp dp
(8.7.6)
158 Elastoplastic Problems of Spheres and Cylinders [Ch. 8
If the axial force is due to internal pressure only, then F* = P/2 and
€z
2(1 - !1-)
= [32 - 1 1 TP dp
Jp (8.7.8)
In any case €z is a known constant. From the last of equations (8. 7.3),
(8.7.9)
Substituting this relation into the first two of equations (8.7.3), making use
of the equilibrium and compatibility equations (8.7.2), and integrating results
after some algebraic manipulations in the following solution:
sr = p - -1
-- Jp TP dp + -1- Jp€P dp + ( 1 - -
1) C1
p2 p2
1
1 _ f/-2 1 ..L.
p p2
(8.7.10)
so = P - 'T
1
+ _!_ Jp TP dp + -1 - - (€r + Jp €r dp) + (1 + P\)c1
p2 p2 1 f/-2 1 p
2p - + 2p2JP TP dp + -1--2
s = 2P 'T
1
-11- 1
p
€r + 2p2 c1 (8.7.11)
(8. 7.12)
Therefore,
Pcrlt
2
-1(
= {32P 1 + 'T - {3 2 2 1 J/3 TP dp )
_
1
(8. 7.13)
For P .::; Pcrlt> we have the elastic solution, which agrees with the classical
elastic solution. For P ;:::: Pcrlt a plastic zone will spread out to some radius
Pc· The solution for general strain hardening can be obtained by an iterative
or successive approximation method, as indicated for the sphere.
Sec. 8-7] Plastic Flow in Thick-Walled Tubes 159
(8.7.14)
To relate Er to Ep, the two methods indicated in Section 7.6 may be used. If
the definition
(8. 7.15)
On the other hand, if the definition (7.6.20) is used, then it follows that
(8.7.16)
The two definitions differ by the familiar constant 2/V3 and either one can
be used. Since the definition based on the plastic work increment appears
to be more consistent with the Tresca criterion, we shall use it, and assume
for the case under consideration that (8.7.14) may be written
lSI ;::: 1
(8.7.17)
lSI ::;; 1
(8. 7.18)
S = /(IErl} sgn S
Er = o
lSI;:::
lSI::;;
1}
1
(8.7.19)
Therefore,
2
c1
1
= ,82 - 1
[JP TP dp
1
,8 (1 - m)
+ p - m(l - /h2) (In Pc - Sr,c - P)
] (8.7.21)
c1 =
p2 [ 1 + T(Pc) - p~
2 2 fPc
1 TP dp
] - p (8.7.23)
2 fPc ]
~2 [ 1 + T(p0) - p~ 1
TP dp
=
,82 [
,82 - 1 p + ,82lJP1 TP dp 1-m
- m(l - /h2) (ln Pc - Sr,c - P)
] (8.7.24)
,8~ - 1 [ 2 fPc ]
Sr,c = - 2,8~ 1 + T(Pc) - ,8~ _ 1 Pc TP dp (8.7.25)
Equation (8.7.26) gives the relationship between the plastic zone radius Pc
and the applied pressure P, for a given temperature distribution r. For a
perfectly plastic material this reduces to
P = In Pc - Sr,c
{3~ - 1 [ 2 (f3c ]
=In Pc + 2[3~ 1 + r(pc) - [3~ _ l JPc rp dp (8.7.27)
To obtain the stresses we now substitute into equations (8. 7.1 0). Thus,
since
P€~ 1-m
J 1
- dp
p
= - - (In p - Sr -
m
P) P :o; Pc
P €~ 1- m
= ---,:;:z- (In Pc - Sr,c - P)
J 1
Pdp P :2: Pc
2 2
p - 1 JP pr dp p)
Sr = m(l
l
- 11-
2
) (
C1 - -
2.- -
1
2 - 2
-11-m p p 1 P
1-m
+ 1 2 (In P - P)
-11-m
P :o; Pc
So = m(l -
1 _ 11-2m
!1-2) (c1 P2 + 1 + _!_ JP pr dp + p - r)
P2 P2 1 P2
+ 11-m
-11-m
(1 + ln p 2 - P)
and (8.7.28)
Sr = -2
IJP rp dp + (l 1-m (In Pc 2)
·
- Sr,c - P)
p 1 -11-m
P :2: Pc
So = - 'T + 2p1 JP1 rp dp + (I I-11-m
- m
2) (In Pc - Sr,c - P)
+ ~ + ( 1 + :2)c1
To obtain the complete stress distribution, we compute P or Pc from (8.7.26),
C1 from (8.7.23), Sr,c from (8.7.25), and the stresses from (8.7.28).
162 Elastoplastic Problems of Spheres and Cylinders [Ch. 8
Problems
1. Show that for the sphere with radial symmetry, the von Mises yield criterion
becomes
2. Explain why one would expect the Tresca and von Mises yield criteria to
coincide for the case of a sphere with radial symmetry.
3. Derive equations (8.2.1) and (8.2.3).
4. Obtain the equations for all the strains and displacements in the ·sphere
before yielding begins for pressure loading only, for thermal loading only,
and for the case when both thermal and pressure loading exist.
5. Show that the steady-state temperature distribution in a sphere of inner
radius a and outer radius b is equal to
T = Toa (!?. _ 1)
b- a r
if the inner and outer surfaces are kept at temperatures of T 0 and zero,
respectively.
6. Using equations (8.3.3) and (8.3.8), show that the stresses are continuous
across the elastoplastic boundary.
7. Compute the displacements and strains in a sphere with pressure loading
only, for r :o; rc. Assume a perfectly plastic material and that the dimensions
remain fixed. Determine the error in the displacements of the inner radius
for the fully plastic case if J1, is assumed to equal 0.5 instead of 0.3.
8. Show that for a hollow sphere with a temperature distribution given by
equation (8.2. 7), the tangential stress is compressive and the radial stress
is near zero in the region adjacent to the inner circumference, so that the
yield criterion in this region can be written
Se- S, = -1
9. Starting with equations (8.6.1), derive equations (8.6.2) and (8.6.4) using
boundary conditions (8.6.6).
10. Derive equation (8.6.9).
11. Derive equations (8.6.12).
12. Derive equation (8.7.6).
13. Derive equations (8.7.10) through (8.7.12).
14. Show that the definition (7.6.20) for the equivalent plastic strain increment
leads to equation (8.7.16) for the case of a tube with the Tresca criterion
and associated flow rule, if a 0 > a 2 > a,.
General References 163
15. Plot the plastic zone as a function of the applied pressure for a sphere with
linear strain hardening. Assume f3 = 2, m = 0.1, and 1'- = 0.3. Compare
the results with those for a perfectly plastic material.
16. Repeat Problem 15 for a tube.
17. Perform complete numerical analysis of the problem of the sphere,
a :$ r :$ b, T(a) = T0 , and T(b) = 0. Assume E = 30 x 10 6 , 1'- = 0.3,
a= 10- 5 , a.= 30,000 + 136,000 (ep + 10- 4 ) for ae > 30,000, and E, a,
and 1'- are independent of temperature.
References
1. W. Johnson and P. B. Mellor, Plasticity for Mechanical Engineers, Van
Nostrand, Princeton, N.J., 1962.
2. R. Hill, The Mathematical Theory of Plasticity, Oxford Univ. Press, London,
1950.
3. P. G. Hodge and G. N. White, A Quantitative Comparison of Flow and
Deformation Theories of Plasticity, J. Appl. Mech., 17, 1950, pp. 180-184.
4. D. N. de G. Allen and D. G. Sopwith, The Stresses and Strains in a Partially
Plastic Thick Tube Under Internal Pressure and End-Load, Proc. Roy. Soc.
(London), A205, 1951, pp. 69-83.
5. M. C. Steele, Partially Plastic Thick-Walled Cylinder Theory, J. Appl. Mech.,
19, 1952, pp. 133-140.
6. R. Hill, E. H. Lee, and S. J. Tupper, The Theory of Combined Plastic and
Elastic Deformation with Particular Reference to a Thick Tube Under
Internal Pressure, Proc. Roy. Soc. (London), A191, 1947, pp. 278-303.
7. W. T. Koiter, On Partially Plastic Thick-Walled Tubes, Biezeno Anniversary
Volume in Applied Mechanics, N. V. De Technische Uitgeverij H. Starn,
Haarlem, 1953, pp. 232-251.
8. D. R. Bland, Elastoplastic Thick-Walled Tubes of Work-Hardening Material
Subject to Internal and External Pressures and to Temperature Gradients,
J. Mech. Phys. Solids, 4, pp. 209-229.
9. I. S. Tuba, Elastic-Plastic Analysis for Hollow Spherical Media Under
Uniform Radial Loading, J. Franklin Inst., 280, 1965, pp. 343-355.
General References
Hill, R., The Mathematical Theory of Plasticity, Oxford Univ. Press, London,
1950.
Hoffman, 0., and G. Sachs, Introduction to the Theory of Plasticity for Engineers,
McGraw-Hill, New York, 1953.
Johnson, W., and P. M. Mellor, Plasticity for Mechanical Engineers, Van Nost-
rand, Princeton, N.J., 1962.
CHAPTER 9
THE METHOD
OF SUCCESSIVE
ELASTIC
SOLUTIONS
In Chapter 8 it was indicated how the sphere and tube problems can be
solved for arbitrary strain hardening by a successive-approximation method.
This method is nothing more than the extension of Picard's method (see
reference [1]) of successive approximations to nonlinear equations. The
method was apparently first used in plastic flow problems by Ilyushin [2]
in his treatment of a thin shell. Ilyushin refers to it as the method of successive
elastic solutions, since each iteration involves essentially the solution of an
elastic problem.
Before the advent of modern high-speed computing machinery, this
method could be used only for relatively simple problems. However, with
current widespread availability and use of digital computers it now becomes
possible to solve simply and quickly many problems whose elastic solution
can be obtained by numerical methods.
Before proceeding to describe this method for the general elastoplastic
problem, we shall first give an illustration of the method of successive
approximations for a simple differential equation [3]. Consider the equation
dy
- -y = 0 y(O) = 1 (9.1.1)
dx
164
Sec. 9-1] General Description of the Method 165
The solution is known to bey = ex. To find the solution by Picard's method,
we proceed as follows. Integrate (9.1.1) to give
(9.1.2)
y<l) = 1
Substitute this value for y on the right side of (9.1.2) and calculate a second
approximation for y:
y< 2l = 1 + J: y< 1l dx = 1 +x
Substitute the second approximation for y and calculate the third approxi-
mation:
y<3l = 1 + rx y<2l dx =
Jo
1 + X + X2
2
Continue in this way to get
y<4l =
rx
+ Jo y<3l dx = 1 + x + 2! + IT
x2 x3
x2 xn
y<n + l) = 1 +
ix
o
y<n) dx = 1 + X + - + · · · + -
2! n!
As n gets larger and larger, it is seen that the infinite series for ex is approached.
The exact solution can thus. be approached as closely as desired by taking
more and more approximations.
This technique can be directly extended to the general elastoplastic problem
in the following manner. For convenience the pertinent equations given in
previous chapters will be repeated here. The equilibrium and compatibility
equations are independent of the plasticity relations and are given by equa-
tions (3.2.2) and (4.7.2); i.e.,
(9.1.3)
166 The Method of Successive Elastic Solutions [Ch. 9
and
(9.1.4)
2
~ ( _ OBxy + OByz + OBzx) = o ez
oz 8z ox 8y ox oy
The stress-strain relations depend on the plasticity theory used and we can
write
(9.1.5)
1 p A 1-
Byz = 2G Tyz + Byz + UByz
_ 1 p A P
Bzx - 2G Tzx + Bzx + UBzx
where s~, e~, etc., are the total accumulated plastic strains up to, but not
including, the current increment of loading, b..s~, b..s~, b..sf, etc., are the
plastic strain increments due to the current increment of loading. The plastic
strain increments are related to the stresses through the yield criterion and
Sec. 9-1] General Description of the Method 167
the associated flow rule. For definiteness we shall consider the Prandtl-Reuss
relations, but any other set of relations can be used equally well. Thus
(9.1.6)
where
l1Bp =
v2v(l1B~-l1B~) +(l1B~ -1:1BD +(1:1B~ -l1B~) + 6[(l1B~y) +(l1B~z) +( l1B~
2 2 2 2 2
x?J
3
(9.1.7)
A P -
LJ.Bx - -l1Bv (2 Bx' - By' - Bz')
3Bet
A P -
uBy - -l1Bv (2 By' - Bz' - Bx' )
3Bet
(9.1.8)
168 The Method of Successive Elastic Solutions [Ch. 9
where
Finally, the equivalent plastic strain increment llBp is related to the equivalent
stress a. or the equivalent total strain Bet through the stress-strain curve. In
addition, of course, the boundary conditions must always be satisfied. For
complete generality, the full three-dimensional equations have been written
out above. In practice only one- and two-dimensional problems can usually
be solved, which is equally true for elasticity problems.
The method of successive approximations now proceeds as follows. The
loading path is divided into a number of increments. For the first increment
of load, a distribution is assumed for the plastic strain increments llB~, llB~,
etc. The total plastic strains B~, etc., are zero. The set of equations (9.1.3),
(9.1.4), and (9.1.5) are now solved as for any elasticity problem and a first
approximation obtained for the stresses and total strains. At the same time,
using the assumed values of the plastic strain increments, an equivalent
plastic strain increment llBp is computed by the first of equations (9.1.7).
From the stress-strain curve the corresponding value of a. can be determined.
A new approximation can now be obtained for the individual plastic strain
increments using the Prandtl-Reuss relations (9.1.6).
Using these new plastic strain increments, equations (9.1.3) through (9.1.5)
are solved again as a new elastic problem. A second and presumably better
approximation is obtained for the stresses and total strains. At the same time,
using these last values of the plastic strain increments a new approximation
is computed for the equivalent plastic strain increment llBp by the first of
equations (9.1.7). Using this value of llBp, a new value is obtained for a8
from the stress-str.ain curve. New approximations are now obtained for the
plastic strain increments !lBL etc., using the Prandtl-Reuss relations (9.1.6).
The process is continued until convergence is obtained; i.e., the differences
between two successive sets of strain increments are less than some prescribed
values. The calculation scheme is illustrated by the flow diagram of Figure
9.1.1.
In this manner the solution is obtained for the first increment of loading.
For the next increment of load, an exactly similar calculation is made except
that B~, B~, etc., are no longer zero but are equal to the known values of
!lB~, llBP, etc., obtained for the first increment ofloading. The complete stress
Sec. 9-1] General Description of the Method 169
I
-- p
6eij,n f- 6ep f-- ""e ~
p
6elj,n+1 1----
Differential
Equations ""ij 1-
and strain history can thus be obtained for any loading and unloading path.
If a deformation type of theory of plasticity is used, only one loading step
is required for the calculation.
The question of course which immediately arises is: What about conver-
gence? Will this method always converge and, if not, under what conditions
will it converge? A rigorous discussion of the convergence problem for the
complicated set of nonlinear equations representing the elastoplastic problem
is beyond the scope of this book. Our discussion will therefore be based on
experience, together with some qualitative observations.
From the previous description of the method and Figure 9.1.1, it is seen
that the equivalent plastic strain increment !:iep is computed from the individ-
ual plastic strains by means of the first of equations (9.1.7), and the equiva-
lent stress a. is then determined from !:ieP and the stress-strain curve. This is
illustrated by path OABC of Figure 9.1.2. On the other hand, it would seem
more direct to compute a. from the second of equations (9.1.7) and then to
determine l:ieP from a. and the stress-strain curve as shown by the flow
diagram of Figure 9.1.3 and path OCBA of Figure 9.1.2. However, experience
has shown that this latter scheme will generally not converge, whereas the
former method will converge provided the loading increment is made sufficiently
small. The reason for this can be seen qualitatively from Figure 9.1.2. Since
the stress-strain curve is very flat in the plastic region, it follows that a small
error in D..eP will produce a smaller error in a., but a small error in a. will
produce a much larger error in D..eP, and so the second method described will
not converge. The reason for the suggested order of the computations for
the sphere and tube of Chapter 8 [equations (8.6.8) and (8.7.18)] now becomes
apparent.
It was mentioned above that the first method described will generally
converge if the increment is made small enough. The question of what
increment size is sufficiently small for a given problem can usually be deter-
mmed only by trial and error; i.e., one picks an increment size and if the
process diverges the increment size is reduced. As a rough rule of thumb it
has been found that if D..eP is less than about 0.3 per cent, convergence will
usually take place, whereas if it is greater than 0.3 per cent, the process may
diverge. The reason for this divergence can again be seen in a qualitative
way from the Prandtl-Reuss relations. Thus consider the equations (9.1.6),
In using this equation the value of D..eP is computed from the previous values
of D..e~, D..e~, etc., using the first of equations (9.1.7). The value of a. is then
obtained from the stress-· train curve, using this computed value of D..eP, as
shown by path OABC of Figure 9.1.2. The stresses appearing in this equation
are obtained from the solution of equations (9.1.3) through (9. 1.5), using the
previous values of the plastic strain increments, as indicated by Figure 9.1.1.
If we compute a. by the second of equations (9.1.7) using these stresses, the
value of a. will not in general agree with the value of a. obtained from the
stress-strain curve, until convergence is obtained. There is thus an incon-
sistency in the Prandtl-Reuss equations (9.1.6), this inconsistency diminishing
as convergence is approached. If the increments of strain are too large, this
initial inconsistency will produce divergence.
In the second method described above and shown in Figure 9.1.3, this
type of inconsistency is avoided, since a. is computed from the calculated
stresses and D..eP is then obtained from the stress-strain curve. However, as
Sec. 9-1] General Description of the Method 171
pointed out above, the second method diverges because of the flatness of the
stress-strain curve.
Both these difficulties can, however, be avoided by using the plastic
strain-total strain equations described in Section 7.9 and given by equations
(9.1.8) and (9.1.9). In using these equations the total strains are obtained
from the solution of equations (9.1.3) through (9:1.5) with assumed or
previously calculated approximate values of tle~, tle~, lle%, etc. The modified
or shifted total strains e~, e~, e~, etc., are then computed from the first of
equations (9.1.9) and the equivalent total strain from the last of equations
(9.1.9). The equivalent plastic strain increment is then determined from eet
and the stress-strain curve using the relation (7.9.16); i.e.,
(9.1.10)
where ae,t- 1 is the value of ae at the end of the previous increment ofloading,
and, if we are dealing with the first increment, ae,t _ 1 is equal to the yield
stress. The new values of the plastic strain increments are then computed
from equations (9.1.8), and the process is continued until convergence is
obtained. A flow diagram of the calculation scheme is shown in Figure 9.1.4.
~ toe f.
1;,n
H H H H H
D. E. G" ••
I}
e,j- let toe
p
H{. r
FIGURE 9.1.4 Block diagram for rapidly convergent scheme.
In this case, the values of e~, e~, e~, etc., appearing in equations (9.1.8) are
consistent with the value of eet appearing in these equations. Also, examination
of (9.1.1 0) shows that since the denominator of the right side is approximately
unity, a plot of !lev versus eet will have a slope of approximately unity, so that
a small error in eet will produce the same order-of-magnitude error in !lev,
without any magnification.
As a result, it has been found from experience that this last method is
convergent for even large size increments and converges more rapidly than
the first method described when that method converges. The plastic strain-
total strain method is therefore generally recommended for use wherever
possible.
We shall now present a series of examples of the use of the successive
approximation method. These examples are taken from references [4] and [5]
and also unpublished results. In all the examples of this chapter the von Mises
criterion and the Prandtl-Reuss equations will be used.
172 The Method of Successive Elastic Solutions [Ch. 9
FIGURE 9.2.1 Thin infinite strip with temperature distribution across the width.
-
ex - E1 Gx + a T + exP (9.2.1)
aT p e~ y
T =- Ex=- 7]=- (9.2.2)
eo eo c
Ex= H
s +T+ Ep (9.2.3)
(9.2.4)
which gives
(9.2.5)
Sec. 9-2] Thin Flat Plate 173
l p
J-1
SdrJ =-
aoch
=P*
(9.2.6)
(9.2.7)
where
(9.2.8)
- (A 2 - A 3 'r)) [f 1
H(r + Ep)'r) d'rj + M*] (9.2.9)
174 The Method of Successive Elastic Solutions [Ch. 9
E = (A 1 - A 27]) (f 1
Hr d7] + P*)
- (A 2 - A 3 7])(f 1
Hr7] d7] + M*)
+ (A 1 - A 27]) f 1
HEp d7] - (A 2 - A 3 7]) f 1
HEp'l) d7]
-T (9.2.10)
+ 2I J1-1 Ep d7] + 23 7) J1
-1 Ep'l) d'l) (9.2.11)
(9.2.12)
these equations are linear Fredholm equations of the second kind and the
solution can often be obtained in closed form.
60 X10 3
i -- /_
c-ep
~
I
1/
0 0.004 0.008 0.012 0.016 0.020
Strain e
FIGURE 9.2.2 Typical stress-strain curve.
X 10- 3 X 10- 3 X 1Q-3 X 10- 3 x 10- 3 X 10- 3 X 10- 3 x 1o- 3 X 10- 3 X 10- 3
1 0 0 1.9 0.7 1.7 0.5 1.6 0.4 1.6 0.4 1.6 0.4
2 0.1 0 1.8 0.6 1.6 0.4 1.5 0.4 1.5 0.4 1.5 0.4
3 0.2 0 1.7 0.5 1.5 0.4 1.4 0.3 1.4 0.3 1.4 0.3
4 0.3 0 1.4 0.3 1.2 0.2 1.1 0.1 1.1 0.1 1.1 0.1
5 0.4 0 1.0 0 0.8 0 0.7 0 0.7 0 0.7 0
6 0.5 0 0.5 0 0.3 0 0.2 0 0.2 0 0.2 0
7 0.6 0 -0.2 0 -0.4 0 -0.4 0 -0.5 0 -0.5 0
8 0.7 0 -0.9 0 -1.1 -0.1 -1.2 -0.2 -1.2 -0.2 -1.2 -0.2
9 0.8 0 -1.7 -0.5 -1.9 -0.7 -2.0 -0.8 -2.0 -0.8 -2.0 -0.8
10 0.9 0 -2.7 -1.5 -2.9 -1.6 -3.0 -1.7 -3.0 -1.7 -3.0 -1.7
11 1.0 0 -3.8 -2.5 -4.0 -2.7 -4.0 -2.7 -4.1 -2.8 -4.1 -2.8
(9.2.14)
By using Figure 9.2.2 and the procedure outlined above, Table 9.2.1 was
constructed. Eleven equally spaced stations were used and the integral was
evaluated by the trapezoidal rule. If greater accuracy is desired, a greater
number of stations and a more accurate integration formula can be used.
However, the accuracy of Table 9.2.1 is sufficient for engineering purposes,
as can be seen in Figures 9.2.3 and 9.2.4, where the results of Table 9.2.1 as
well as the results obtained by using 51 stations and Simpson's rule are
4.0 X10- 3
--elastic
- - - f i r s t approximation
3.0 - second and third
approximations
2.0
;;;t:::-- .........
1.0
"-":::.;
~
.,~
~
0
~
~~
"'
c
0 '~)
t; -1.0
~\_
~.
-2.0
~~
\\~[\
-3.0
~~
-4.0
\
-5.0
0 0.2 0.4 0.6 0.8 1.0 ,,
Distance from plate center line
FIGURE 9.2.3 Strain distribution in infinite strip.
178 The Method of Successive Elastic Solutions [Ch. 9
100 X10
--elastic
- ---first approximation
- - second and third
60
==1'---r--.,.......
approximations
20
r-- - ~
·v; '~ ~
': 0
~
b" -20 ~,,
lil
~
~ r- ,_
'\
ti 1\.
-60
\
-100
1\
-140
0 0.2 0.4 0.6 0.8 1.0
Distance from plate center line
FIGURE 9.2.4 Stress distribution in infinite strip.
For linear strain hardening the solution can often be obtained in closed
form. We shall consider only the simple type of temperature distribution of
the previous example. A more general formulation is given in reference [5].
Thus let T = T 0 (ry 2 - t). To will be called the loading parameter. For linear
strain hardening it can readily be shown that (assuming constant material
properties)
Ep = (1 - m)(E + 1) (9.2.15)
(9.2.16)
Sec. 9-2] Thin Flat Plate 179
where
K(''), g)= 0
K(TJ, g) = 1
If 1€1 < 1 everywhere, no plastic flow will occur and the first term on the
right of (9.2.16) gives the elastic solution. The maximum value of 1€1 will
obviously occur at TJ = 1 and yielding will therefore begin at the edge,
TJ = 1, when E = -1, i.e., in compression. The value of To for yielding to
begin is then
'To,crlt = -!
If To < t, no yielding will occur. As the loading parameter To is increased
beyond t, a compressive plastic zone will spread inward from the edge. The
center of the plate, TJ = 0, will meanwhile be in a state of tension, the tensile
stress increasing as To increases. Eventually at some value of To designated
by To,c a tensile plastic zone will begin spreading from the center toward the
compressive zone spreading from the edge.
Let YJ 1 be the value of TJ at the edge of the tensile plastic zone extending
from the center, and let YJ 2 be the value of TJ at the edge of the compressive
zone spreading from the edge; i.e., for 0 < TJ < YJ 1 there is a tensile plastic
zone with € ;:::: 1, and for YJ 2 ::; TJ ::; 1 there is a compressive plastic zone with
E ::; -1. As long as To ::; '~"o,c• YJ 1 = 0, and equation (9.2.16) becomes
1
€ = -To(TJ 2 - t) + (1 - m) r (€ + 1)dg
Jn2
1
-To(TJ 2 - t) + (1- m)(1- ")2) +(1-m) r Edg
Jn2
(9.2.17)
Jn2
€(
1) dg
- '~"o(TJ 2 - t) + (1 - m)(1 - TJ 2 ) + (1 - m)
x [;o (TJ~- ") 2) +(I - m)(1 - YJ 2) 2]
+ (1 - m) 3 (1 - ") 2 )3 + (1 - m) ;o (TJ~ - YJ 2 )
x [1 + (1 - m)(1 - "12)]
180 The Method of Successive Elastic Solutions [Ch. 9
+ (1 - m)2(1 - 77 2) 2]
or
n
e<n+ 1l = -To(71 2 - j-) + (1 - m)(1 - 77 2) L (1 -
!=0
m)f(1 - 712)1
n-1
- m) ~ (71~ - 712) ~ (1 - m)f(l - 712) 1
0
+ (1
1
and
lim e<n+1) = e = -To(7l2 _ j-) + (1 _ m) 1 - 712 + (To/3)(71~ - 77 2)
n->oo 1 - (1 - m)(1 - 772)
(9.2.18)
The above derivations of equations (9.2.18) and (9.2.20) are actually not
necessary, since the integral equations (9 .2.19) and (9 .2.17) are Fredholm
equations of the second kind with degenerate kernels and the solutions can
therefore be obtained directly by standard methods (see, for example,
reference [6]). However, the successive approximation method was used in
obtaining the solution to show how this method leads to the exact solution
for this more complicated case.
Equations (9.2.18) and (9.2.20) give the complete solution to the problem
if 77 1, 77 2, and To,c were known. These can be determined as follows. If
To :::; To,c, there is one region of plastic flow extending from 77 2 to 1. At
Sec. 9-2] Thin Flat Plate 181
Thus for given values of m and To, 'TJ 2 can be determined or, more simply,
for given values of m and 'r) 2 , To can be calculated directly. For a perfectly
plastic material, m '= 0, and
'TJ2--
- ( 3 )l/3 (9.2.22)
2T 0
Once 'r) 2 is known, the complete strain distribution is obtained from (9.2.18),
the plastic strains and the stresses from (9.2.15) and (9.2.3).
If To is greater than To,c, a tensile plastic zone will spread from the center
to 'rJl· To determine the values of 'r) 1 and 'r) 2 in this case, one first sets 'rJ = 'TJ 1
and € = 1 in (9.2.20) and then 'rJ = 'r) 2 and E = -1 in (9.2.20). This results
in the following two simultaneous equations for determining 'TJ 1 and 'TJ 2 :
=
1 -To('r)t + j-) + C
(9.2.23)
-1 = -To('r)~ + j-) + C
where
C = (1 _ m) 1 - 'TJ1 - 'TJ2 + iTo('TJ~ - 'TJ2 - 'f)~ + 'r) 1)
1 - (1 - m)(1 + 'TJ1 - 'TJ2)
The value of To,c at which the center begins to flow plastically can now be
found. For this condition, 'TJ 1 = 0 and (9.2.24) gives
'TJ2=[f (9.2.25)
To,c
Substituting this value for 'TJ 2 together with 'rJl = 0 into either of equations
(9.2.23) results in
2 2
Tg c +61 - 2m To c +9 (1 - 2m) To c - 32 (1-- -
m) = 0 (9.2.26)
• m · m • m
Only one of the roots of(9.2.26) will be physically meaningful. For a perfectly
plastic material, m = 0, and
32
To,c = 9 = 3.555 (9.2.27)
182 The Method of Successive Elastic Solutions [Ch. 9
Figure 9.2.5 shows the variation of To,c with m. The value of To,c is rather
insensitive to the strain-hardening parameter m.
1'6~
0 3.4
0.
tJ>
If To is greater than To,c, two regions of plastic flow will exist. The corre-
sponding values of 71 1 and 71 2 are found by solving simultaneously (9.2.24) and
one of (9.2.23). Thus eliminating 71 1 from the first of (9.2.23) by means of
(9.2.24) results in the following quintic for 71 2 :
5
712 +
( 3m
4(1 - m) +2
1-
mT0
m) 712 + 3m1 ( m + To3 ) 712
4 3
1.0
0.8
0.6
"7
0.4
0.2
2 4 6 8 10 20 40 60 80100
Loading parameter To
The next problem we shall consider is that of a thin circular shell with an
axial temperature gradient. We shall use the deformation theory of plasticity
for this problem and assume a one-step thermal load, although this is not
essential.
The equilibrium equations for the shell are given in reference [7]. These are
Nx = 0
Rd M 2
(9.3.1)
[2 dx2 x + Ne = 0
h/2
where Nx =
J -h/2
h/2
ax dz
N8 =
J -h/2
h/2
a8 dz (9.3.2)
Mx
J
=
-h/2
axzdz
where R is the mean radius of the shell and l is the characteristic length,
equal to
(9.3.3)
ax = - E 2 [ex - aT-
1 -tL
e~ + fL(ee - aT- e~)]
(9.3.4)
ae = - E 2 [ee - aT -
1 -tL
e~ + fL(Bx - aT - e~)]
(9.3.5)
w
ee = - R
184 The Method of Successive Elastic Solutions [Ch. 9
where u is the axial displacement of a point on the middle surface of the shell
and w is the radial displacement, positive inward. Substituting (9.3.5) into
(9.3.4) gives
ldu w + (1
dx = t-t R + tt)aT + 7i1 J"'
-n/
2
2
(e~ + tte~)dz (9.3.7)
(9.3.8)
w
ex = t-t R + (I + tt)aT + 7i1 J"' 2(e~ + ttendz -
2
e2 = -1-t-t -+-f1-aT - -
-f1-- ( ex+ ee) +11-t-t 1 --2t-t
- (ex+
p
e8P) (9.3.10)
1-t-t
Finally from the second of (9.3.1), making use of (9.3.8), there results
(9.3.11)
Sec. 9-3] Thin Circular Shell 185
12[2 fh/2
where P = -h
3 (e~ + f.Le:)z dz
-h/2
(9.3.12)
4R fh/2
Q = -h e~ dz
-h/2
(9.3.17)
186 The Method of Successive Elastic Solutions [Ch. 9
28 X10- 3
24 /
....
o;"'
.~ 20 /
....~ /
"'
016 /
....0 /
+-
-::: 12 /
!!!
0
/
>
·::; 8 /
rr
w /
4
/
0 4 8 12 16 20 24X10- 3
Equivalent plastic strain Bp
Then from (9.3.3), l = 3.81 in. and 0 ::; x ::; 12.6. Assume a temperature
distribution given by T = 2.21x 2 (corresponding to a 350°F rise from one
end of the shell to the other end), and boundary conditions (fixed ends)
(e~ + 0.3ef)z dz
1
P = 21.8 J -1
(9.3.19)
Q = 24 f 1
-1
e[ dz
Also
g) dg
(9.3.21)
Twenty-one stations were taken along the length of the shell and six stations
through the thickness. For the zeroth approximation, it is assumed that
e~ and e: and hence 11 and / 2 are zero. The function w(x) is calculated from
(9.3.20) with the constants C2 and C4 determined from the last two boundary
conditions of equation (9.3.18). The strains ex, e0 , and e 2 are then computed
from equations (9.3.5), (9.3.9), and (9.3.10). First approximations to e~ and
e: are obtained by computing Bet from the third of equations (9. 3.17), reading
ev from Figure 9. 3.1 and calculating e~ and e[ from the first two of equations
(9.3.17). With these values of e~ and e:, P and Q are computed from equations
(9.3.19), 11 and 12 from (9.3.21), and w from (9.3.20) with new values com-
puted for c2 and c4. New values are then computed for Bx, eo, and Bz and the
process continued until convergence is obtained.
188 The Method of Successive Elastic Solutions [Ch. 9
(9.3.22)
Approximation
--
---
0
1
-- 4 through 7
40 X10 3
20 II r\
1{1
~4
0
·u;
~
a.
~
+-
~ -80
.Q
><
<(
-100
-120
-140
-160
0 2 4 6 8 10 12 14
Dimensionless axial distance x
(a)
0
...,_ ~ ~
·v;
a.
f.- Approximation
\
be;,
~ -20 I- --o l\
.g ---1
f.- - - 4 through 7 \ ~-
:J
V>
<ll -40 \ '\'
"0
·u;
+-
:J
\
0
b -60
~
~
t;
0 -80
~
<ll
0\
c
~-100
-120
I
-140
0 2 4 6 8 10 12 14
Dimensionless axial distance x
(b)
I I
Approximation
2.0 x10- 3 - I--
--o
---1
- - 4 through 7
.4~
"tl
.,><
I
1.0
!Jl
t 0 ~l'
.g
:J
V>
<ll
] -1.0
+-
:J
0
c
~ -2.0
·et;
]~ -30 .
-4.0
\
-5.0
0 2 4 6 8 10 12 14
Dimensionless axial· distance x
(c)
190 The Method of Successive Elastic Solutions [Ch. 9
2.0 x1Q- 3
Approximation
h.
t:l
--o
I
Q)
---1
1.0
"' - - 4 through 7
"'
-,
(.)
0
't:
:::>
If) 0
Q)
'"0
·u;
+-
:::>
0 -1.0 ~\
c
0
c
Th
·e ~
t; -2.0
0
~ ~
Q)
Ol
c
-3.0
\
~
-4.0
'
0 2 4 6 8 10 12 14
Dimensionless axial distance x
(d)
and
~~ = ! sin(x - g) sinh(x - e)
d 3G
dx3 = cos(x - e) cosh(x - e)
All the integrals in the previous equations were evaluated using the trape-
zoidal rule. The results for this problem are shown in Figure 9.3.2. An
abbreviated calculation setup for one of the successive approximations at the
last two stations is given in Table 9.3.1. As can be seen from the figures, the
differences between the seventh and the fourth approximations are very small.
From an engineering viewpoint the first approximation is actually sufficient.
This is probably due to the fact that the total strains do not change much,
because of the plastic flow. This will generally be true when the only loads
are thermal loads.
Sec. 9-2] Thin Circular Shell 191
Q p /1 I•
Sta- Bx- aT e6 - ixT ePX e: w
tion (9.3.9) (9.3.5) (9.3.t7) (9.3.17) (9.3.t9) (9.3.2t) (9.3.20)
l J 1 6 Jo-
l r8-
b- -l- -<.2-
~- t-8- -g- -2l- t-o-ood- 8 r-- 8- _()_
N 0 ~
0 0
0
eeP
0
0
0
0
l{),--10
0 ~ l{) l{)
4X 110-3 N
f--r- I -I 0 0 0
0 o_ 0 c:5 0
c:5 c:5
0
c:5
I I I l -,o.o~25
2 L-- I
lL
'(.--"
l-
--
/
l
-
........-;
/
7
v I-
-r--
..__
\ r--4..
r- I I
0.0?15
J.
r--
....... Vf I -r ..-
v r--
\ ~ \
o.r101_
v/ I / ~r-....~ ~ \ \ I 0.0005
-- f.- ~
/r
J 1; ' .\ ~ - .1
I
0~00b1
0 exP
0.0001_
-I
~.0005-
-
0 1/
\ :---.. sf\. lfJ- ___..... t-
I I
I r-\ \ t" r- -~ v I /'1 I I
:..- ~.00110-
\ "\- ......... i / / f1' I l/
vI
-2 I 1-- -\._ / ....1-
.-1v
0.~01~-
I
--
I - I- l£- I--
\ I
-; 1--
t--
- f--.
1
I_ _., t--
I 1 1-'
I
f-""' 0.0:020 -
1
I
l j
l 1/ /
-4
- [...--
l-lo.oo25
I I
t--
-4 -2 2
(9.3.24)
Sec. 9-4] Long Solid Cylinder 193
(9.3.25)
(9.3.26)
89
_
a
T _
-
(I + 2(23E- p,) ~) 8p
P
89
+ 2(1 - 2p,) ~
3E 8p
P
8x
(9.4.1)
d8 9 89O - 8r
-
dr + - -
r =
(9.4.2)
194 The Method of Successive Elastic Solutions [Ch. 9
II= p,E E
(1 + p,)(1 - 2p,) G = 2(1 + p,)
To save writing, er and e: are here defined as the total accumulated plastic
strains, including those for the current increment of load; i.e.,
1-1
er = 2 b.ef.k + b.er,l
k=1
(9.4.3a)
1-1
e: = 2 b.e:,k + b.e:,l
k=1
where the summations are assumed to have already been computed and
b.e:.i and b.ef. 1 are to be determined. Assuming E, p,, and ez constant, and
substituting (9.4.3) into (9.4.1) and eliminating er by the use of (9.4.2), a
differential equation is obtained for e0 , which upon integration results in
For a solid cylinder C 2 must vanish. Now substituting equation (9.4.4) into
the compatibility relation (9.4.2) and solving for e, gives
To determine ez use is made of the fact that ez is a constant and that for
unloaded ends
(9.4.6)
Substituting into the third of the stress-strain relations (9.4.3) and solving
for e2 results in
1 - 3p, 1 fR 1 - 2p, fR ef - eC
C1 = 1 _ f.t R 2 Jo aTr dr + 2( 1 _p,) Jo r dr
+ 2(1
1 1
- p,) R2 JorR (e, + eoP)r dr
p
(9.4.8)
e0 = -1-+-f.t21 aTrdr
1 -p,r o
J,' + 21( 1- -p,
2p,)2
r
1 J,' (e, + e )rdr
o
p p
0
1 - 2p, f' ef - eC
+ 2(1 - p,) Jo -,.- dr + cl
(9.4.9)
-eo
1 + p, T
+ -1--
-p,
1 - 2p, p 1 - 2p,
a + -1-- e, + -1--
-p, -p,
f' o
ef - eC d
r
r + 2c 1
prev.
.p
r
·~
prev.
.,
Eq. Eq.
•e
Eq.
••
Eq. Fig.
Bet Bp eP
r
Eq.
•f
Eq.
n r aT approx. approx. (9.4.9) (9.4.9) (9.4.9) (9.1.9) 9.3.1 (9.1.8) (9.1.8)
These values of t.er and t.eC are substituted into equations (9.4.8) and (9.4.9)
and new approximations obtained for e" e8 , and e2 • The process is repeated
as many times as necessary to obtain the desired degree of convergence. After
1100
l;'- 1000
~ \.
::J
+-
2 900
\
Q)
\
a.
E
~ 800 1\
\
700
0 0.2 0.4 0.6 0.8 1.0
Radius, in.
FIGURE 9.4.1 Temperature distribution in long solid cylinder.
Approximation
0.010 0
----- 1
- - 4ond5
~ \
0.009
~
"'\; ee
1\
0.008
\
\
"'0.007
c
~
0.006
\\
0.005
\'\
\ er
\\
0.004
\
0.003
0 0.2 0.4 0.6 0.8 1.0
Radius, in.
{a)
FIGURE 9.4.2 Strain and stress distributions in long solid cylinder.
Sec. 9-5] Rotating Disk with Temperature Gradient 197
- - 4 and 5
/
80
(
1/
60
·' (]"8-
I
I
f_,.-
20
0
I <rr
convergence, the stresses can be computed from (9.4.3). The load is then
incremented and the process repeated.
The above calculations have been carried out for a 1-in.-radius cylinder
with a temperature distribution as shown in Figure 9.4.1, using the strain-
strain curve of Figure 9.3.1. The computations are shown in Table 9.4.1 for
one iteration and the results are plotted in Figure 9.4.2. Very little difference
is found for this problem between the deformation theory and the incre-
mental theory assuming the temperature distribution applied in several steps.
the strains similar to the cylinder problem of the previous section. The
second method converts the differential equations to finite-difference form
and solves the resulting finite-difference equations. The integral-equation
formulation has the advantages of being conceptually simpler and of including
the boundary conditions automatically. On the other hand, it is difficult by
this approach to take into account variations of material properties with
temperature as well as variations in disk thickness. These can readily be
taken into account by the finite-difference method.
Integral-Equation Formulation
(9.5.1)
where h is the disk thickness, p the density, and w the rotational speed. For
the plane stress problem a 2 = 0, and the stress-strain relations can be written
[see equations (9.3.4)]
E
ar = - - -2 [e,
1 -p,
+ p,ee - (er + p,ef) - (1 + p,)aT]
(9.5.3)
ae = - E [ee
1 -p,2
+ p,e, - (e~ + p,ef) - (1 + p,)aT]
The terms er and e~ are here defined, as for the cylinder, by equations (9.4.3a).
Proceeding as for the case of the cylinder in Section 9.4, equations (9.5.3)
are substituted into (9.5.1) and e, eliminated by the use of (9.5.2), to give a
differential equation in ee. Assuming that h, E, and p, are constant, this
differential equation is integrated to give
1 - p, 2 pw 2r 2 1 + p, J' 1- p, Jr eP - e~
ee =- - E - --
8
+ --r2
-
o
aTrdr + - -
2 o
r
r
dr
1 + t-t
+-y,:z J' (e, +eer d. +y+f2
0
p P) 1 cl c2 (9 ,5,4)
TABLE 9.5.1 Calculation of Plastic Strains in Rotating Disk.
eP
r e: er ee ez eet ep e: e:
prev. prev. Eq. Eq. Eq. Eq. Fig. Eq. Eq.
n r aT approx. approx. (9.5.5) (9.5.4) (9.5.7) (9.1.9) 9.3.1 (9.1.8) (9.1.8)
x 10-s x 10- 3 x 10- 3 X 10- 3 x 10- 3 x 10- 3 x 10- 3 X 10-s x 10- 3 x w- 3
1 0 0.95 1.567 1.567 3.757 3.757 -3.247 4.669 3.441 1.720 1.720
2 0.5 1.045 1.515 1.517 3.780 3.784 -3.033 4.543 3.317 1.657 1.660
3 1.0 1.33 1.459 1.351 3.983 3.845 -2.490 4.270 3.048 1.573 1.474
4 1.5 1.71 1.376 1.144 4.250 3.942 -1.776 3.919 2.721 1.466 1.252
5 2.0 2.47 1.256 0.7269 4.862 4.104 -0.3885 3.277 2.124 1.298 0.8067
6 2.5 3.80 1.089 0.1111 6.022 4.381 1.913 2.389 1.304 1.047 0.1508
7 3.0 5.415 1.070 -0.4288 7.614 4.797 4.371 2.035 0.9972 0.9899 -0.3908
8 3.5 7.315 1.410 -1.175 9.773 5.362 6.964 2.578 1.469 1.371 -1.142
9 4.0 9.5 1.904 -2.298 12.28 6.076 10.00 3.623 2.446 1.908 -2.279
10 4.5 11.88 2.290 -3.637 14.83 6.911 13.51 4.896 3.664 2.303 -3.620
11 5.0 14.25 2.498 -4.996 17.18 7.824 17.18 6.235 4.982 2.491 -4.982
-
~
(Q
(Q
200 The Method of Successive Elastic Solutions [Ch. 9
I - fl-2 pw2r2
-eo- - E - - - +(I + p,)aT + ef + p,e~
2
i
r eP eP
+ (I - p,) r -
0
dr + C1 (9.5.5)
o r
(9.5.6)
to give
(9.5.7)
Finally, the constant C1 is evaluated from the known rim loading. If the rim
stress due to the rim loading is ar(R), the first of equations (9.5.3) gives
(R being the radius of the disk)
ar(R) = --p,
EI2 [er + p,e 0 - (ef + p,en - (I + p,)aT]r=R (9.5.8)
and substituting equations (9.5.4) and (9.5.5) into (9.5.8) gives for C1 :
- 2 J0
I rR ep -
r r
ep
0
dr
I
+ 2R 2 J0
rR ( ef + ef)r dr
] (9.5.9)
1600
v
lL.
0
i
::J
1200
/
v
b..... 800
(l)
0. /
E
~ 400 /
0
!'-"-" - ... v 2
Radius, in.
3 4 5
Approximation
0
---- 1
-- 4
--- 12 with intermediate
extrapolation or 40
without extrapolation
0.018
er
0.016 I
II
0.014
//
II
0.012 VI
/
IJ
'<)
.s
~ 0.010
+-
(/) ,~
,J/ eg
0.008
) /
~f
v
/, /~
0.006
Lf1 v_ v
-~- li':P" ----- ~
~ k-;:::' ~
.- --
~
0.004
·-
0.002
0 1.0 2.0 3.0 4.0 5.0
Radius, in.
(a)
Approximation
0
-----1
-.---4
- - - - - 1 2 with intermediate
extrapolation or 40th
without extrapolation
120 X10 3
~ ~ r-..
...............
80
40
- -= f::.::::-
\
t---
F.};::. ;-\ ,,
·~
"""'
~
k\' ~- ~ ......_......_
\~ f-:::::,_~
0
~
.~
"r
·v; -40
0.
vl
V1
\ '~
\ ""'"" "e
~-
-
~
+-
(J)
-80
\
1\
-100 \
1\
-140
\
\
-180
0 1.0 2.0 3.0 4.0 5.0
Radius, in.
(b)
5.2 X10- 3
0 _Br
~\
o 8e
5.0 \ \
\
4.8
\
', \
ll
\
·~ 4.6
+-
({)
' \. \
'\.
\
4.4
\ \
I~
4.2
'\
~
4.0
jq
0 0.04 0.08
Se
FIGURE 9.5.3 Variation of strain with change of strain.
Finite-Difference Formulation
where ef and e~ represent the total plastic strains up to the current increment
of loading, and ~ef and ~e~ are the increments of plastic strain due to the
current increment of loading.
Substituting (9.5.10) into the compatibility equation (9.5.2) gives the
compatibility equation in terms of stresses:
Equations (9.5.1) and (9.5.11) are two equations for the two stresses ar and a8.
We proceed by putting them into finite-difference form as follows. Let the
disk radius be divided into N intervals (not necessarily equal). There are thus
N + 1 stations, the first station being at the center for a solid disk, or at the
inner radius for a hollow disk. The last station is at the outer radius. Equa-
tions (9.5.1) and (9.5.11) are written in finite-difference form at the midpoints
of these intervals. Thus at the midpoint of the (i - 1)st interval,
(9.5.12)
Sec. 9-5] Rotating Disk with Temperature Gradient 205
Fil - f-1-i-1 -
---
bi
Ei-1
2
Hi = ~ (ri - ri-1)(pihirt + Pi-1hi-1rt-1)
(9.5.14)
where a1, a1_1> L~o and M1 are the indicated matrices. Equation (9.5.17)
represents a linear recurrence relation.between the stresses at the ith station
and the stresses at the (i - l)st station. Obviously by successive application
of (9.5.17), the stresses at the ith station can be linearly related to the stresses
at the first station. Let this linear relation be written
(9.5.18)
where A1 and B1 are as yet unknown and a 1 are the radial and tangential
stresses at the first station. Substituting (9.5.18) into (9.5.17) gives
A1 = A1-1L1
(9.5.20)
Hence (9.5.21)
Beginning therefore with A 2 and B2 as given by (9.5.21), all the other A's
and B's can be computed successively by the recurrence relations (9.5.20).
For the last station [the (N + l)st], rN+l equals R, the disk radius, and
equation (9.5.18) becomes
or
ar,N+l] = [all,N+l a12,N+l] [ar,l] + [bl,N+l] (9.5.22)
[ae,N+l a21,N+l a22,N+l ae,l b2,N+l
Sec. 9-5] Rotating Disk with Temperature Gradient 207
Therefore,
ar(R) - b1 N+l
Ur,l = ae,l = ' (9.5.23)
au,N+l + a12,N+l
For a disk with a central hole of radius R 0 , with inner prescribed pressure
we get
Ur,l = ar(Ro),
(9.5.24)
Thus a 1 is now known. The stresses at every station can then be directly
computed by means of (9.5.18).
To summarize then, the L1 and M1 matrices are computed from (9.5.16)
and (9.5.14), the A1 and B1 matrices from (9.5.20) and (9.5.21), and then the
stresses are computed from (9.5.18), using either (9.5.23) or (9.5.24).
It is to be noted that this straightforward procedure takes into account
with equal ease variations of E, h, p, or even f1- along the radius of the disk.
Furthermore, if the dimensions of the disk are changing during the plastic
flow process, this can readily be taken into account. For if r is the current
radius to a given point P, and r' was the radius to the point P before plastic
flow took place, then approximately
H
(9.5.26)
h = (1 + ef)(l + e&}
Thus at any stage of the plastic flow process the values of r1 and h1 appearing
in equations (9.5.14) can be corrected by means of (9.5.25) and (9.5.26).
The finite-difference formulation presented will of course give directly and
quickly the elastic solution for a disk of arbitrary profile with variable
properties, if P{ is set equal to zero in equations (9.5.16). For the plastic
problem P[ is not zero, and its values can be determined by successive
approximations, as thoroughly described in the previous examples. We shall
208 The Method of Successive Elastic Solutions [Ch. 9
(9.6.1)
deo eo - er O
-
dr + - -r =
(9.6.2)
(9.6.5)
Sec. 9-6] Circular Hole in Uniformly Stressed Infinite Plate 209
Integrating results in
The triple integral can be somewhat simplified by making use of (9.6.5) and
integrating by parts. The constants A and B are determined from the bound-
ary conditions
(9.6.7)
A=~ (9.6.8)
2
and the constant C can arbitrarily be set equal to zero without affecting the
solution.
Detailed calculations were carried out by Tuba [10] using the successive-
approximation technique with linear strain hardening for various values of
the strain-hardening parameter m (ratio of slope of the strain-hardening part
of curve to slope of the elastic part), and for various ratios of a oo to the yield
stress a 0 • His results, taken from reference [10], are shown in Figures 9.6.1
through 9.6.3. In these figures the stress and strain concentration factors due
m=O
m=0.1
3.0
m=0.2
m =0.3
Kee m=0.4
m=0.6
m =0.8
2.0 m =1.0 (elastic)
m=0.8
m=0.6
m=0.4
Ko-e m =0.3
m=0.2
m=0.1
1.0 m=O
0 1.0
m
K = e8(a) = e8(a)
Be e8 (oo) [(I - p..)/E]a"'
K _ ee(a) _ e.(a)
Be - e.(oo) - i[(I + p..)/E]a"'
Problems
-dy + y
dx
= 0 y(O) = 1
References
THE PLANE
ELASTOPLASTIC
PROBLEM
213
214 The Plane E1astop1astic Problem [Ch. 10
(10. 1.1)
(10.1.2)
(10.1.3)
(10.1.4)
A
uBy
p
= -~Bp ( 2B y - I I
B2 -
I )
Bx
3Bet
B~ = L ~e~,k
k=l
Sec. 10-1] General Relations 215
e~ = ex - B~ etc.
(10.1.6)
B~t = 3v2 [( I
Bx -
1)2
By +( I
By -
1)2
Bz
+ (BzI -
1)2
Bx
+ 6(Bxy
I )2]1/2
(a 2
ae-- X + a Y2 - a XY
a + 3r 2XY ) 1 ' 2
(10.1.7)
(10.1.8)
and
(10.1.10)
(10.1.13)
plane strain
(10.1.14)
) = _E_ (8 ( Lle~) + 8 ( Lle~) _ 2 8 ( Lle~y))
2 2 2
Ll (x
g ' Y - 1 - ft 2
8y 2 28x ox 8y
- ~.tE V1 2 ( Lle~ + Llet)
1 - { t2
The equilibrium equations (10.1.1) are identically satisfied and the com-
patibility equation (10.1.11) becomes
(10.1.16)
The boundary conditions to be satisfied by the stress function rp are (see, for
example, reference [1])
(10.1.17)
rp =
J (orp
s
-
dx
-
ox ds
+ -orp -dy) ds = f(s) + Ca
oy ds
(10.1.18)
drp = orp dy _ orp dx = h(s)
dn ox ds oy ds
so that if orp(s)joy and orp(s)joy are known from (10.1.17), orpjdn and rp(s)
can be computed from (10.1.18). For an unloaded boundary, Tx = Ty = 0
and consequently j 1(s) = j 2 (s) = 0, so that orp(s)joy = C1 and orp(s)jox = C2 ,
and, since the stresses depend only on the second derivatives of rp, we can
218 The Plane Elastoplastic Problem [Ch. 10
r
2b '-------~x
1!-----~
~---------2o--------~~
external loads, the boundary conditions (10.1.19) are used; i.e., f and its
normal derivative are zero on the boundaries.
Equation (10.1.16) with the right side zero is the classical biharmonic
equation. The elastic thermal stress problem with g = D..g = 0 has been
Sec. 10-2] Elastoplastic Thermal Problem for a Finite Plate 219
•
t; /+2
•
/-1,)+1
•
i,/+1
• it1,/+1
8
• • • • •
i-2,/ ;-1, j t;) i+1,j 1+2, j
•
i-1,/-1
•
i, j-1
•
it1, /-1
•
i, j-2
1
84 [r/>t-2,, + r/>1.1-2 + r/>1.1+2 + r/>1+2,,
+ 2(</>t-1,1-1 + <Pt-1,1+1 + 4>t+1,j-1 + 4>t+1,1+1)
+ 20r/>t,1- 8(</>t-1,1 + 4>t,1-1 + 4>t.1+1 + 4>t+1,1)]
= - E'v 2 ( aT) 1,1 - g1,1 - ll.g1,1 (10.2.1)
algebraic equations. Once these are solved, the stresses can be computed
from the relations
Txy,!j =
cfl-1,}+1 - cf!-1,1-1 -
4S2
cfi+1,J+1 + cfl+1,1-1
The strains are computed from (10.1.3) and the plastic strain increments
from either (10.1.4) or (10.1.5), together with the stress-strain curve. The
function 11g is now changed and the solution obtained again. The process is
continued until convergence is obtained. It is to be noted that only the right
sides of the set of n x m simultaneous equations change from iteration to
iteration. It should also be noted that although equation (10.2.1) has been
written for equal spacing between stations, it is possible to write the finite-
difference formulas for unequal spacing.
Once the calculation has converged for a given increment of load (in this
case, thermal load) and /1g determined, /1g is added to g, the load is incre-
mented, and a new calculation started to determine the value of the plastic
strain increments and stresses due to the new increment of load. It is to be
remembered that at any station for which ae is less than the yield stress, the
plastic strain increments are set equal to zero.
As an example, such a solution was obtained for a square plate with a
parabolic temperature distribution given by T = T0 (y 2 - f), the constant T0
being raised in increments until large zones of plastic flow occurred. Linear
strain hardening was assumed with the strain-hardening parameter m taken
to be 0.1. Some of the results are shown in Figures 10.2.3 and 10.2.4. Only
one quadrant of the plate is shown, the other three quadrants being identical
because of symmetry. In these computations 20 stations were taken in the
x direction and 20 in the y direction, resulting in a set of 400 simultaneous
equations to solve.
Figure 10.2.3 shows the rate of growth of the regions of incipient flow
The curves are the loci of all the points of incipient plastic flow for a ·
value of loading parameter To = T 0 Erxja 0 • Plastic flow starts first at
centers of the four sides of the plate and moves rapidly inward. Plastic
does not start at the center of the plate until it is well developed at the
Once plastic flow has started at the center, however, the rate of growth
this zone is greater than at the sides.
Figure 10.2.4 shows the plastic strain trajectories for the maximum
Sec. 10-2] Elastoplastic Thermal Problem for a Finite Plate 221
1.0
/ /
v //
0.9 V/
//
v ./
0.8
To ... v v/
~-
I
v / v
0.7
~;.... v 1/ /
0.6
i-
-- ~ !--"" .......-i><-15
f- 20 To v
/
vv
/ V_ v
y 0.5 2b
-
20
0.4
== =-1F-
I r-
r-..... 157¥ v 11
I-
........
1'- I I /
.'\ f/
-
~
0.3
= ~1
1\ II I
0.2 r-... \ \ I
\ I I
0.1
\ \ I ij
1\ \ I
1_ l 1 I
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
X
1.0
r-g k-
..-
......... v _,...v v /_ ~ '/-
r- 7_ f.-
..-
v . . . v v ~ r.%'V I I I I I
0.9
J- f- "5
0.8 F= F= -4 /
~ /
6~
k-
1---::
- ~ t% ~
/
/
/': b( V' r'---0.5
"/
v v h Vj~f/
v
./
- -incipient
plastic
flow
~ -:l ~ % 1'-1
0.7 r-
f.- 1- ...... --
~ ......... )(::: v/ "'--1.5
L ~
0.6
J- f-
J- 1- - ~ ~/ "-2
Dimensionless
plastic strain,
Ep / ~
I%- !1-j
1.% ~ v
vv
y 0.5
I ''. '// '/ II"1/
t--t--, lfJ IJ J J J
0.4
r-- r-._0.5
"" \ 1.5:-' fL '/ v vv v
r- 1-- 1"'- Ill rn l_; L L /_
0.3
- I-r-.... '\ I\ 0.5_::' fL 1/ 1/ 1/ I!
0.2
!"'--
: , 2. 2\
1.5 "
~
1 \
\ \
I
II I I
1--c h( I ,12/3 ~ I I
I I ,,
0.1 I'-- I lit II I 415 /6
'i
......... A<
~~2.4
2.5
\ I I I 1/
~~ \ I \ 1 J J _j ~
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
X
€
e -aT
= ....::X::.._--.-=,-
Uo/E
1.2 2
infinite strip
0 3x1plate
- ..............
0.8 1\ <=
·~ 1
\ t;
"'vi'
0.4 i
\
VI
VI
~ 0
.2
VI
""' '\ \
VI c
~
+-
VI 0 1\ ,g -1 1\.
I~
'()
VI
VI ~
0 1\
~
c
·~ -0.4
\
(
+-
~ -2 \
c
<))
E
0
Q.
E
1\
0 -0.8 \ .o
u -3 \
\ 0
,',!
c
\
-1.2
"-...," ~
0
-4 \
['., <))
:0::
-1.6 -5
0 0.2 0.4 0.6 0.8 i.O 0 0.2 0.4 0.6 0.8 1.0
Dimensionless coordinate, y
(a) (b)
Since the solution obtained was an incremental one using finite increments
of load, it was thought worthwhile to run an experiment to determine the
effect of increment size f1r 0 on the final solution. It was found that f1r 0 = 2.5
was approximately the largest increment size that produced no appreciable
difference in the final stress and strain distribution at r 0 = 20. Calculations
were also performed using the two iterative techniques described in Section
9.1, that is, using equations (10.1.4) to calculate the plastic strain increments
and using equations (10.1.5). As expected for very large load increments, the
first method (using the Prandtl-Reuss relations) diverged, whereas the second
method, using the plastic strain-total strain equations, always converged.
For those cases where both methods converged they gave identical answers.
The solution to this problem can be found discussed in greater detail in
reference [2].
The above calculations were performed for a very thin plate. For a very
thick plate, a condition of plane strain would exist at planes far from the
surfaces. To obtain a solution to this problem only minor modifications of
previous calculation method are necessary. Thus g and f1g must be computed
using equations (I 0.1.14) instead of (10.1.13). Also it is now necessary to
compute a 2 • This can be done as for the case of the cylinder in Section 9.4.
From the third of equations (10.1.8),
(10.2.3)
L a2 dA = 0 (10.2.4)
The elastic solution of the infinite plate with a crack was first obtained by
Inglis [3] and was used by Griffith [4] in his theory of brittle crack propagation.
It was pointed out by Orowan [5] that, for ductile materials, the plastic strain
energy was a major factor in the energy balance of the system and could not
be ignored in any analysis of crack stability. Since no solution was available
224 The Plane Elastoplastic Problem [Ch. 10
for the plastic strain field, various assumptions were introduced to take into
account the plastic zone at the tip of a crack, such as Irwin's "equivalent
crack length" [6] and Neuber's "plastic particle" [7].
In the present section it will be shown how the previous method can be
used to obtain at least approximately the elastoplastic strain field in the
vicinity of a crack, for a strain-hardening material. Since the solution can be
obtained only numerically, the presence of the mathematical singularity at
the tip of the crack makes accurate answers very difficult to obtain. The
solution to be presented is therefore not intended to provide accurate quanti-
tative results, nor is it necessarily, or even probably, the best method of
solving this problem. It is intended primarily to provide qualitative informa-
tion on the effect of strain hardening and on the differences that might be
expected between plane stress and plane strain solutions.
Consider the case of an infinite plate with a central crack 2 units long, with
a uniaxial tensile load at infinity perpendicular to the plane of the crack.
As before, all the stresses are made dimensionless by dividing by the yield
stress and all the strains are divided by the yield strain. The tensile stress at
infinity is also divided by the yield stress. Since half the crack length is taken
as unity, the x andy coordinates are dimensionless in terms of the half-crack
length. We shall attempt a solution using finite differences, as in the case of
the rectangular plate of Section 10.2.
We are faced, however, with two problems in applying the finite-difference
methods previously described. In the first place, it is obviously impossible to
cover an infinite region with a finite-difference grid. Second, the crack tip is a
singular point of the stress field and, as pointed out in reference [8], the error
in the finite-difference formulation in the vicinity of the singularity spreads
to other points. The best procedure in this type of problem is to subtract out
the singularity, if possible, giving a new problem with different boundary
conditions, but which will be well behaved.
In the present problem we attempt to minimize both the above difficulties
by subtracting out the elastic solution from the problem. The elastic solution
contains a singularity at the crack tip and also satisfies the boundary condi-
tions. We are thus left with a well-behaved problem with homogeneous
boundary conditions. Furthermore, it has been shown experimentally by
Dixon [9] that for most materials (mild steel, which has a lower yield point,
is one exception), the strain field outside the plastic zone is the same as the
elastic strain field; i.e., the elastic solution prevails outside the plastic
The assumption is therefore made that the infinite plate can be replaced by
finite rectangle with an edge crack upon whose boundaries the elastic
field acts. Because of symmetry, only one quadrant of the plate need
considered. We proceed, therefore, in the following manner.
Sec. 10-3] Elastoplastic Problem of the Infinite Plate with a Crack 225
~I _l __
-1
X
infinite plate acts. The lower boundary is a line of symmetry. The differential
equation to be satisfied is equation (10.1.16) without the temperature term;
i.e.,
V44> = -g(x, y) - ~g(x, y) (10.3.1)
Let
(10.3.2)
where 4>e is the elastic solution to the problem. Then 4>e satisfies the differential
equation
(10.3.3)
(10.3.4)
</>v, satisfying equation (10.3.4), symmetric about the lower boundary and
equal to zero along the crack and on the other three boundaries of the quad-
rant. This function can be found using finite-difference and successive approxi-
mations just as was done for the thermal stress problem of Section 10.2.
The rectangle shown is covered by a grid and the differential equation
(10.3.4) is written in finite-difference form for every point of the grid. The
solution then proceeds by successive approximations as previously described.
The second iterative method employing the plastic strain-total strain equa-
tions was used. One additional step is, however, needed. Mter the function
<Pv is obtained by solving the set of simultaneous equations, the elastic stress
function, cp 8 , must be added to <l>v before computing the total stresses. The
rest of the iterative scheme proceeds exactly as before. Both the plane stress
and plane strain solutions can be readily obtained, as described in Section 10.2.
The elastic stress function </>e has been obtained in many ways. Using
Muskhelishvili's solution [10], it can be written
where Re stands for the "real part of," z is the complex variable x + iy, z is
its conjugate, and P is the stress at infinity divided by the yield stress. For
every iteration the value of cp 8 , as computed from equation (10.3.5), is added
to the <Pv values at every station before the stresses are computed.
The first computation made was to determine the validity of the assumption
that the elastic stress field prevails outside the plastic zone. For this purpose
three different-sized rectangles were chosen to enclose the plastic zone, as
shown in Figure 10.3.2. For each of these rectangles a plane stress calculation
was made by the method described to determine the extent of the plastic zone.
Linear strain hardening was assumed with the strain-hardening parameter m
taken equal to 0.1, and the load was raised in steps in four increments from
a value of 0.2 times the yield stress to 0.5 times the yield stress. The elasto·
plastic boundaries at a load of 0.5 times the yield stress are shown in the
figure. It is seen that there is not too great a difference in the plastic zone size
for all three rectangular boundaries used, particularly the last two, indicating
that the effect of the plastic zone on the stress field outside the zone dies out
rapidly, as indicated by Dixon's experiments [9].
Also shown in Figure 10.3.2 is the elastic yield locus, i.e., the locus of
points which are just at the yield stress as computed elastically. This
commonly assumed to be the boundary of the plastic zone. It can be seen
from the figure that this assumption can be appreciably in error. This
further illustrated in Figure 10.3.3, where the elastoplastic boundary is
Sec. 10-3] Elastoplastic Problem of the Infinite Plate with a Crack 227
o small grid
t::. medium grid
o large grid
I r------------------1 !
0 I I I
II :
' II ot::. II
[p o oo ~:::.
I
0
II ;srJJJ- . . . . , I '
t \,
0 D
y
Ii
0
°;"& i I
0
I II :1~: :. I
~ tt I
I
1
00
I
'i \mokt;p / oH i '
0.8 0.9 1.0 1.1 1.2 1.3 1.4
X
1.0
0 1-----'----::0.
0.8 0.9 1.1 1.2
X
FIGURE 10.3.4 Growth of plastic zone size with load; plane stress.
P=0.7
::;;..-_(,..
- _u_ _ _ _ ,___ _L _ _ _ j_
FIGURE 10.3.5 Growth of plastic zone size with load; plane strain.
were solved for one particular case, less than 300 of the grid points were
within the plastic zone. It is, of course, possible to improve this situation by
using a variable grid spacing, which was not done in the above calculations.
Second, the effect of the singularity on the results of the computation is
not at all clear. The elastic solution cf>., which is subtracted as previously
described, has a square-root stress singularity. If the plasticity solution has
the same order singularity, then the technique used in subtracting the singu-
larity would be effective. However, it is likely that the order of singularity
changes as plastic flow progresses. This has been indicated in results obtained
by Swedlow [11]. If this is true, then subtracting the elastic singularity may
not be so effective and errors due to the singularity may be propagated
through the solution. This requires further investigation.
A solution using the successive-approximation technique and finite differ-
ences was also obtained in reference [12] for a finite plate with symmetric
edge cracks. Again, because of the coarseness of the grid and the stress
singularities, the results cannot be expected to be accurate, but do give an
indication of the shape and growth of the plastic zone.
230 The Plane Elastoplastic Problem [Ch. 10
(10.4.1)
ev can then be obtained from the plot of ev versus eet derived from the
stress-strain curve. The plastic strains are computed from
(10.4.2)
It should be noted that the total strains, being a solution of the elastic
problem, satisfy the compatibility relations. The stresses, however, do not
satisfy equilibrium equations. If the plastic zone is relatively small compared
to the elastic zone, then, since the elastic stresses do satisfy the equilibrium
equations, equilibrium will be satisfied approximately on the whole, and the
strain-invariance answers would be expected to be reasonably correct.
However, if the plastic zone is very large, then equilibrium on the whole
will not be satisfied and the strain-invariance answers may be appreciably in
error.
The results of strain-in variance answers for the plate problems of Sections
9.2 and 10.2 are shown in Figures 10.4.1 and 10.4.2. Figure 10.4.1 shows a
---strain-hardening error
- - - - strain-invariance error
- · - error due ta neglecting
strain-hardening and
assuming strain- invariance
'
\
\
\ 0.1_.----
---
,,,........
\\
.........
.,.,.. .....
'· __ _.5::--------
....
Loading parameter r0
10
___ p at e
------ --
neglecting strain hardening x
3 1 1
or-----~--------~~~-=----------------------
\
'-----
---·
neglecting strain hardening
- · - 1x1 plate
10 20
Loading parameter r 0
plot of the error in the maximum plastic strain assuming strain invariance,
as a function of the loading parameter r 0 , for the thin infinite strip of Section
9.2. Also shown is the error in neglecting strain hardening for a strain-
hardening parameter m = 0.1 (i.e., assuming m = 0 instead of m = 0.1).
It is seen that the error in neglecting strain hardening is on the conservative
side (gives larger plastic strains than actually exist); the error in strain
invariance is on the nonconservative side. It is interesting to note also, as is
shown in the figure, that if one makes both assumptions, i.e., neglects strain
hardening and assumes strain invariance, the errors tend to cancel each other.
In any case, from an engineering viewpoint a 10 per cent or even a 20 per cent
error is not very large, considering our inexact knowledge of material
behavior. Figure 10.4.2 shows similar results for the square plate and the
3 x 1 plate of Section 10.2.
Problems
1. Derive equation (10.1.11).
2. Obtain the finite-difference equation corresponding to equation (10.2.1)
assuming the grid spacing in the y direction is twice that in the x direction.
3. Determine the expression for e. for the case of generalized plane strain by
means of equations (10.2.3) and (10.2.4).
4. Show that the stress function given by equation (10.3.5) satisfies the boundary
conditions for an infinite plate with a crack 2 units long and loaded at infinity
with a tensile stress P normal to the plane of the crack.
References 233
References
1. I. S. Sokolinkoff, Mathematical Theory of Elasticity, McGraw-Hill, New
York, 1956, p. 260.
2. E. Roberts, Jr., and A. Mendelson, Analysis of Plastic Thermal Stresses
and Strains in Finite Thin Plate, NASA TN D-2206, 1964.
3. C. E. Inglis, Stresses in a Plate Due to the Presence of Cracks and Sharp
Corners, Trans. Inst. Naval Arch., 55, No. 1, 1913, pp. 219-239.
4. A. A. Griffith, The Phenomena of Rupture and Flow in Solids, Phil. Trans.
Roy. Soc. London, A221, 1921, pp. 163-198.
5. E. Orowan, Fundamentals of Brittle Behavior in Metals, Fatigue and Fracture
of Metals, M.I.T. Press, Boston, Mass., and John Wiley, New York, 1952,
pp. 139-169.
6. Fracture Testing of High Strength Sheet Materials, A Report of a Special
ASTM Committee, ASTM Bull. 243, 1960.
7. H. Neuber, Theory of Notch Stresses, ABC TR-4547, 1958.
8. L. Fox, Numerical Solution of Ordinary and Partial Differential Equations,
Pergamon Press, London, 1962, p. 301.
9. J. R. Dixon, The Effect of Local Plastic Deformation on the Stress Distri-
bution Around a Crack, NEL Rpt. No. 71, Dec. 1962.
N. I. Muskhelishvili, Some Basic Problems in the Mathematical Theory
of Elasticity, P. Noordhoff, Groningen, 1953.
J. L. Swedlow, M. L. Williams, and W. H. Yang, Elasto-Plastic Stresses
and Strains in Cracked Plates, Proceedings of the International Conference on
Fracture, Sendai, Japan, Sept. 1965; also Galcit SM 65-14, July 1965.
I. S. Tuba, A Method of Elastic-Plastic Plane Stress and Strain Analysis,
J. Strain Anal., 1, No. 2, 1966, pp. 115-120.
A. Mendelson and S. S. Manson, Practical Solution of Plastic Deformation
Problems in the Elastic-Plastic Range, NASA TR R-28, 1959.
S. S. Manson, Thermal Stress and Low Cycle Fatigue, McGraw-Hill, New
York, 1966, p. 196.
M. J. Turner, R. W. Clough, H. C. Martin, and L. J. Tupp, Stiffness and
Deflection Analysis of Complex Structures, J. Aerospace, Sci., 23, No. 9,
1965, pp. 805-823.
CHAPTER 11
THE TORSION
PROBLEM
z
''f--'r-------
1 ,'
I ,
For the case of the beam twisted as shown, it is assumed that the angle of
twist of a given cross section is directly proportional to its distance from the
origin. Thus if the angle of twist per unit length is a, then a section a distance
z from the origin will have rotated an amount
1{1 = az (11.1.1)
Now consider a point P originally at location (x, y, z). This point will
rotate to a position P 1 with coordinates (xi> Y1> z1 ), where
x1 = x +u JI=y+v z1 = z +w
u, v, and w being the displacements, assumed small. The projection of these
points on the xy plane are shown in Figure 11.1.2. If the polar coordinates
of the point P are r and 8, then we have
x = r cos 8 y = r sin 8
(11.1.2)
x1 = rcos(8 + 1{1) Y1 = r sin (8 + 1{1)
then
or u = -yza
v = xza (11.1.4)
and w = w(x, y, a)
Bxz
. 1(
= 2 - ya + 0:
8 )
(11.1.5)
Byz =
1(Xa + 8w)
2 oy
~ (- Oeyz + Oexz) = O
ox ox oy
_ ~ ( _ Oeyz + Oexz) = O
oy ox ay
8s os
or _J!3. - ___3.!, = constant
ox ay
Sec. 11-1] Torsion of Prismatic Bar. General Relations 237
From (11.1.5) it follows that the constant in (I 1.1. 7) must equal a. Therefore,
the compatibility equation for this problem becomes
0Byz 0Bxz
---=a (11.1.8)
ox By
It should be emphasized again that the shear strains eyz and exz are the
components of the strain tensor as defined by equation (4.1.12) and are there-
fore equal to one half the engineering shearing strains used by other authors.
Looking now at the equilibrium equations (3.2.2), it is seen by virtue of
(11.1.5) and the stress-strain relations that the first two are identically satisfied
and the third becomes
(11.1.9)
(11.1.10)
where e~2 and e~2 are the accumulated plastic components of the total shear
strains. Substituting (11.1.10) into (11.1.8), the compatibility equation is
obtained in terms of the stresses:
OTyz
- - - - = 2Ga
OTxz + g (X, y) (11.1.11)
ox oy
X
g ( ,y
) = 2G (8e~
ay
2 _ 8efz)
ox (11.1.12)
Then the equilibrium equation (11.1.9) is identically satisfied and the com-
patibility equation (11.1.11) becomes
(11.1.14)
dy
I = cos (n, x) = ds
(11.1.16)
dx
m = cos (n, y) = - ds
or 4> equals a constant along the boundary. In the case of simply connected
boundaries, e.g., solid bars, this constant can be chosen arbitrarily, since we
are interested only in the derivatives of 4>. It is, therefore, for convenience
chosen to be zero, so that the boundary condition becomes
(11.1.19)
or r = jgrad </>I
The resultant shear stress is thus equal to the gradient of the stress function <fo.
Let us now calculate the resultant forces and moments acting on any cross
section. The force in the x direction is given by
where the double integral is taken over the area of the cross section and
~(x, A) and <fo(x, B) are the values of</> at two opposite points of the boundary
at a given value of x, and are consequently equal to zero because of the
boundary condition (11.1.18). In a similar fashion, it follows that the y
component of the resultant stress is zero. Finally, the torque acting on the
section is computed from
M = JJ(ryzX- TxzY)dx dy
(11.1.22)
240 The Torsion Problem [Ch. 11
(11.
Sec. 11-2] Elasticity Solutions 241
(11.2.3)
Hence
(11.2.4)
(11.2.5)
(11.2.6)
(11.2.7)
The maximum stress will occur on the boundary at the point closest to the
axis of the bar at x = 0, y = b. Thus
(11.2.8)
To obtain the warping function w(x, y, a), we first compute the strains from
(11.1.10):
(11.2.9)
242 The Torsion Problem [Ch. 11
(11.2.10)
Therefore,
b2- a2
w = b2 + a2 axy + constant
and since at the origin w must vanish, the constant must equal zero, and
b2- a2
w(x, y, a) = b 2 + a2 axy (11.2.11)
X=-
a
y'j
(11.2.13)
Sec. 11-2] Elasticity Solutions 243
C=-Ga
V3
2a
Hence
f = -Ga (v. r3 x
2a
- a) ( y - - x
V3 - -
3
2a)(y + -=
x 2a)
V3 + -3 (11.2.14)
is the solution to the problem. The stresses, the torque, and the warping
function can be computed as before.
As a final example, consider a bar of rectangular cross section, as shown in
Figure 11.2.2. For this case the solution is not as simple as for the previous
y
a
X
m
0
• • • 0
• •
• • • • • • •
• • • • • • •
(t;J)
• • • • • • • • • j
• • • • • • • • • /-1
• • • •
0
• •
• • • • • • • • •
• • • • • • • • • 3
• • • • • • • • 2
2 3 i -1 n
where h is the grid spacing assumed constant and is the same in both the
x andy directions.
An equation such as (11.2.15) can be written for each of the n x m grid
points, resulting in a set of n x m simultaneous linear equations for the
unknown values of 4> at each of the points. Once the fs are determined as
the solution of this set of equations, the shear stress can be obtained from
equations (11.1.13) by numerical differentiation and the torque for a given
angle of twist per unit length a from equation (11.1.22) by numerical inte-
gration.
Actually the number of equations to be solved is (m - 1) x (n - 1) rather
than m x n, since the boundary conditions require 4> to be zero at the
Sec. 11-3] Membrane Analogy 245
and right boundaries of the quadrant. For a square section, only the points
on the diagonal and to the right of the diagonal (or left) need be considered,
resulting in -!-n(n - 1) equations. Along the lower boundary, because of
symmetry, equation (11.2.15) becomes
(11.2.16b)
(11.2.16c)
Finite
Ref. [2] Difference
It is seen that the solution with this many grid points is sufficiently accurate.
pressure is applied to one side, the membrane will deflect by an amount given
by the solution of the following equations (reference [2], p. 269):
(11.3.1)
If the applied torque is sufficiently large, plastic flow will occur. Since the
maximum stress will always occur at the boundary (see Problem 2 and
reference [5]), a plastic zone will start at some point on the boundary and
spread toward the interior as the torque is increased. Additional plastic
zones may subsequently start at other points in the cross section. For the
case of the torsion problem, the yield criteria of von Mises and of Tresca
both reduce to
where k is the yield stress in simple shear. According to the von Mises
criterion, k is equal to a 0 /V3, and according to the Tresca criterion it is
equal to a 0 /2, where a 0 is the yield stress in simple tension. If the material is
perfectly plastic, then equation (11.4.1) must hold everywhere in the plastic
region. In terms of the stress function <{>,
Sec. 11-4] Elastoplastic Torsion. Perfect Plasticity 247
and cfo = 0 on the boundary of the bar. The elastoplastic boundary is unknown
and is determined from the conditions that cfo and its first derivatives (the
shear stresses) are continuous across this boundary and that the resultant
shear stress is less than or equal to k inside the elastic region. Equation
(11.4.2) can also be written
inside the plastic region. In other words, the slope of the cfo surface is a
constant, equal to k, in the plastic region, it is not greater than kin the elastic
region, and the height and slope of the cfo surface are continuous across the
elastoplastic boundary.
The above conditions on cfo for a perfectly plastic material suggest the
extension of the membrane analogy to a partially plastic bar [6, 7]. A roof
of constant slope, proportional to k, is erected with the membrane as its base.
As the membrane is pressurized and deflects, it will approach the roof. The
region of the membrane corresponding to the region of the bar flowing
plastically will be pressed against the roof and will have the same slope as the
roof. The rest of the membrane, corresponding to the elastic region, will not
be touching the roof and will have a smaller slope.
The membrane-roof analogy furnishes a simple physical and intuitive
picture of the growth of the plastic zones as the torque is increased. To obtain
quantitative results will usually entail a considerable amount of labor. For
the case of complete plastic yielding, the solution becomes much simpler.
In this case the membrane will be in contact with the whole roof, and it is
no longer necessary to use a membrane. Instead, one constructs a roof of the
proper slope. This can be done by simply heaping dry sand onto a plate whose
shape is similar to the cross section of the bar. Since the torque is equal to
twice the volume of sand (see Problem 5), the torque required to produce
complete yielding can readily be determined. Thus for a circle of radius a,
the volume of the sand hill (in this case, a cone) is
where his the height of the heap. Since the slope of the sand hill corresponds
to the shear yield stress k,
h
k=-a
248 The Torsion Problem [Ch. 11
(11.4.6)
Figure 11.4.1 shows the sand-hill analogies for the above two cases.
FIGURE 11.4.1 Sand-hill analogies for circular and rectangular cross sections.
U=-r/>-
- 2Ge0 a
P-
€y =-
e~z
eo
_ '~'xz 1 + P, _ Tyz 1 + p,
Tx = 2Geo = -ao- T xz TY = 2G
eo
= --
ao
Tyz
~
~p-
= ~ M *- M
= 2Ge0 a3
eo
g = xja 'TJ = yja
Sec. 11-5] Elastoplastic Torsion with Strain Hardening 249
where e0 and a 0 are the yield strain and yield stress, respectively, related to
each other by a 0 = Ee0 and a is a characteristic linear dimension of the
cross section.
The system of equations to be solved for a simply connected cross section
can now be written
(11.5.2)
(11.5.3)
U=O on boundary
=
au (11.5.4)
Ty
- ag
(11.5.5)
2 .;-2--2
Et = VJ V Ey + Ex (11.5.6)
(11.5.7)
(11.5.8)
The elastic solution for a bar with a rectangular cross section by means of
finite differences was presented in Section 11.2. To obtain the elastoplastic
solution the function g(x, y) is subtracted from the right side of equation
(11.2.15), which in terms of the dimensionless quantities defined in (11.5.1)
becomes
where
(11.6.2)
2.0
1.6
)(
0
E
...
0 2 3 4
{3
FIGURE 11.6.1 Variation of Tmax with {3.
Sec. 11-6] Bar with Rectangular Cross Section 251
2.4
X
0
n_E
"'1.6
0 2 3 4
{3
where the strain-hardening parameter m is the ratio of the slope of the linear
hardening curve to the slope of the elastic curve, as previously defined. For
the perfectly plastic case m is equal to 0, and for the elastic case m is equal to 1.
The figures show the effects of the strain-hardening parameter and the angle
of twist on the maximum stress, the maximum plastic strain, the size of the
plastic zone, and the torque. The results are also summarized in Table 11.6.1.
Although the calculations are described using deformation theory, a
similar calculation, increasing a in steps, gave almost identical results. This
is in agreement with similar calculations in reference [11]. As shown in
reference [12], incremental and deformation theories give identical results for
a perfectly plastic material of any cross section or a strain-hardening material
of circular cross seption. For strain-hardening materials of noncircular cross
sections they will yield different results. It appears, however, that the differ-
ences will in general be slight.
For a bar with a circular cross section the solution is greatly simplified.
In particular, for the case of linear strain hardening, a closed-form solution
can be obtained. In polar coordinates the displacements are
Ur = 0 u8 = arz (11.7.1)
(11. 7.2)
(11.7.3)
(11. 7.4)
(11.7.5)
(11.7.6)
254 The Torsion Problem [Ch. 11
Let
S6 r
f1e
2Geo
-
= a=p
(11.7.7)
where a is the radius of the bar. Then equation (11.7.6) can be written in
dimensionless form as
(11.7.8)
(11.7.9)
(11.7.10)
2(1 + p,)
Pc = V3 f3
which depends only on Poisson's ratio and the yield strain but not on
stress-strain curve. The value of f3 at which plastic flow just starts is
from equation (11.7.11) by setting Pc equal to 1. Thus the critical value
will be
f3 - 2(1 + p,)
"- v3
Sec. 11-7] Bar with Circular Cross Section 255
(11.7.13)
1 a0
or aG=--
o V3 a
(11. 7.14)
Note that the critical value of p is obtained when the numerator of (11. 7.15)
vanishes, which results again in equation (11.7.11).
Once the equivalent plastic strain is known from equation (11.7.15), the
shear strain and the stress are computed from (11.7.5) and (11.7.3).
have thus obtained a complete solution in closed form. To compute the
define To = Tozf2Ge 0 • Then
M *- M s -2
= 2-G
s0a
- 7r
o
ToP 2dp il (11. 7.16)
tf3p P:::; Po
To = { 21[3P- P
€o P ~Po
256 The Torsion Problem [Ch. 11
results in
Note that for Pc = 1 (no plastic flow) the torque reduces to the elastic torque
as given in equation (11.2.12). For a perfectly plastic material,
m =0 A= ~3 B = -i(l + t-t)
or
3v'3
3
M = 21ra a 0 [l _12v'3a
1 (Ga)3] 3
a0 (11.7.18)
20
18
16
14
12
.810
m
0.20
0 4 8 12 16
{3 I
.I
FIGURE 11.7.3 Variation of M* with (3 for various m. I
I I
258 The Torsion Problem [Ch. 11
Problems
1. Show that the loading is radial for the torsion problem of a bar with a circular
cross section, so that total plasticity theories may be used, as long as there is
no unloading.
2. Show that the maximum shear stress for a solid bar of elliptic cross section
under torsion occurs on the boundary at the point closest to the axis of the
bar.
3. Determine the stresses, the torque, and the warping function for the triangular
cross section bar of Figure 11.2.1.
4. Show that if the lateral surface of the bar is stress-free, the resultant shear
stress must be tangent to the boundary.
5. Show that if pjS is equal to 2Ga, the applied torque acting on a bar is equal
to twice the volume between a membrane of the same shape as the cross
section of the bar and the z = 0 plane.
6. Determine the torque acting on a bar of circular cross section by calculating
the volume under a membrane of the same shape.
7. Show that for the torsion problem the yield criteria of von Mises and Tresca
both reduce to equation (11.4.1).
8. Using the Saint-Venant assumptions (11.1.4), show that for a bar of circular
cross section the radial and tangential displacements become
llr =0 Uo = arz
and consequently the only nonzero strain is e92 = tra.
9. Calculate the torques required to produce a twist of 0.004 rad/in. in a 2-in.-
diameter shaft if the material is perfectly plastic, and if the material strain
hardens with m = 0.1. Assume E = 30 x 106 , p, = 0.3, and a0 = 30 x 103 •
References
1. B. Saint-Venant, Memoire sur la torsion des prismes, Mem. A cad. Sci. Math.
Phys., 14, 1856, pp. 233-560.
2. S. Timoshenko and T. N. Goodier, Theory of Elasticity, McGraw-Hill,
New York, 1951, p. 275.
3. V. Kantorovich and V. I. Krylov, Approximate Methods of Higher Analysis,
P. Noordhoff, Groningen, 1958, p. 70.
4. L. Prandtl, Zur Torsion von prismatischen Staeben, Physik. Z., 4, 1
pp. 758-759.
5. I. S. Sokolnikoff, Mathematical Theory of Elasticity, McGraw-Hill, New
York, 1956, p. 117.
6. A. Nadai, Der Beginn des Fliessvorganges in einem tortierten
Angew. Math. Mech., 3, 1923, p. 442-454.
7. A. Nadai, Theory of Flow and Fracture of Solids, Vol. 1, lVHAJJ'""'".LJ""'
New York, 1950.
8. W. Johnson and P. B. Mellor, Plasticity for Mechanical Engineers,
Nostrand, Princeton, N.J., 1962, p. 132.
General References 259
General Reference
Johnson, W. and P. B. Mellor, Plasticity for Mechanical Engineers, Van Nostrand,
Princeton, N.J., 1962.
Prager, W., and P. G. Hodge, Jr., Theory of Perfectly Plastic Solids, Wiley New
York, 1951.
CHAPTER 12
THE SLIP-LINE
FIELD
260
Sec. 12-1] Plane Strain Problem of a Rigid-Perfectly Plastic Material 261
(12.1.5)
where k is yield stress in simple shear and T has been written for brevity
instead of T xy· The equilibrium equations to be satisfied are
(12.1.6)
over part of the boundary, then the stress-strain relations must be used to
relate the stresses to the strains and the problem becomes much more com-
plicated.
The principal stresses in the plastic field can be written in terms of am and k
as follows:
(12.1.7)
(12.1.8)
The next step is to :find the prindpal directions. We define the first principal
direction as the direction of the maximum principal stress. Let cp be the angle
between the first principal direction and the x axis as shown in Figure 12.1.1.
Then from equation (3.3.6) for the prinCipal directions it follows that
2T
tan2</> = - - -
ax- ay
which gives two values of cp differing by 90°. The second principal direction
is taken 90° counterclockwise from the first.
Sec. 12-1] Plane Strain Problem of a Rigid-Perfectly Plastic Material 263
Having determined the principal stresses and directions, the maximum and
minimum shearing stresses and directions can readily be determined. These
shearing stresses act on the planes bisecting the principal directions as
described in Section 3.4. Their values are given by
(12.1.10)
(} = rP - 45°
1
tan 2B = -tan 2rP
cos 2(} = i
(12.1.12)
. 2B- ay- ax
Sill - 2k
At every point in the plastic field, the angle which the maximum shear
direction makes with the x axis is determined by equations (12.1.11) or
(12.1.12). If curves are now drawn in the xy plane such that at every point
of each curve the tangent coincides with one of maximum shear directions,
then two families of curves called shear lines, or slip lines, will be obtained.
Obviously, since the maximum and minimum shear directions at a point are
orthogonal to each other, the two families of slip lines will form an orthogonal
set. These two families of curves will be called the a lines and (3 lines, respec-
tively.
It should be carefully noted that along an a line a is varying and (3 is
constant, and along a (3 line (3 is varying and a is constant. a and (3 are merely
or curvilinear coordinates used to designate the point under
264 The Slip-Line Field [Ch. 12
consideration, just as x and y designate the point. Thus the point P shown in
Figure 12.1.2 can be designated P(xl> y 1 ) or P(a 3 , {3 2).
{3 lines
The normal stresses acting on the maximum shear planes equals the
average of the principal stresses, as shown by equation (3.4.5). Thus the
stresses acting normal and tangential to the a and (3 lines are given by
L---------------------~x
ax=am-ksin2(}
ay = am + k sin 20
-r = kcos 20
From (12.1.14) it is seen that the state of stress can be determined in terms
of two independent quantities, am and B. The equilibrium equations can be
written in terms of these quantities by substituting (12.1.14) into (12.1.6).
Thus
or defining
we can write
8 8
Bx - cos 28 B - sin 28 B =0
ox Bx By
(12.1.16)
8
Bx - sin 28 B + cos 28 B8 = 0
By Bx By
Now the choice of the x andy axes is arbitrary. If we choose the x andy axes
at a given point to coincide with the a and f3 directions at this point, then
B = 0 and
B B
By = Bf3
(12.1.17)
where C1 and C2 are constants. These equations were first derived by Hencky
in 1923, [14].
266 The Slip-Line Field [Ch. 12
dex - dey _ ax - ay
(12.2.1)
dexy - 1'xy
(12.. 2.2)
It is convenient to divide the strains by dt, the increment of time, and write
these equations in terms of velocities. Of course, these equations remain
homogeneous in t, which acts merely as a scaling parameter. Then
(8vxf8x) - (8vyj8y) _ ax - ay
(8vxf8y) + (8vyj8x) - 27'
(12.2.3)
Now since the principal axes of stress and of plastic strain increment
coincide (see Section 7.2), it follows that the maximum shear stress lines and
maximum shear velocity lines coincide, or that the stress slip lines are the
same as the velocity slip lines. Also the strain rates normal to the IX and {3
directions are equal to the mean strain rates [see equation (4.5.4)]. Therefore,
Sec. 12-2] Velocity Equations 267
Vx = Va COS 8- Vp sin 8
(12.2.5)
Vy = Va sin 8 + Vp COS 8
If 8 = 0, the x axis will coincide with the a direction and the condition that
the normal strain rates be zero can be written
Ba = (8Vx)
ax 0=0
= 0
ep = (8Vy)
8y 0=0
= 0
or, since the x direction is the same as the a direction and the f3 direction is
the same as the y direction,
8v" 88
- - Vp- = 0
8a 8a
(12.2.7)
268 The Slip-Line Field [Ch. 12
These are the compatibility equations for the velocities first derived by
Geiringer in 1930, [15].
If the problem is statically determined, the slip line field and the stresses
can be found from equations (12.1.18) (or their equivalent) and the stress
boundary conditions. The velocities can then be computed from (12.2.8) (or
their equivalent) using the velocity boundary conditions, since d8 will now
be known from the stress solution. If, however, the problem is not statically
determined, which means that the stress boundary conditions are insufficient
to obtain a unique slip-line field, then equations (12.1.18) must be solved
simultaneously with (12.2.8) using both the stress boundary conditions and
the velocity boundary conditions. This is an extremely difficult problem and
must usually be done by trial and error. A slip-line field satisfying all stress
conditions is assumed. The velocities are then computed and a check made
to see if the velocity boundary conditions are satisfied. If not, the slip-line
field is modified and the procedure repeated as often as necessary. This is
obviously a very laborious process, since the construction of just one slip-line
field is a lengthy task.
It is worthwhile to note some of the differences between Hencky's stress
equations (12.1.18) and Geiringer's velocity equations (12.2.8).
1. Hencky's equations relate two unknowns, x and 8, by two equations.
Geiringer's equations relate three unknowns, Vx, Vy, and d8, by two equations.
2. Hencky's equations give the stress state all along a known slip line if the
stress state is known at one point on the slip line. Geiringer's equations will
not give the velocities along a known slip line, if they are known at one point.
3. Hencky's equations force certain restrictions on the geometry of the
slip-line field, as we will shortly see. Geiringer's equations place no restriction
on the geometry of the slip-line field, except through the boundary conditions.
1. Hencky's first theorem states that the angle between two slip lines of
one family at the points where they are cut by a slip line of the other family
is constant along their lengths. This is shown in Figure 12.3.1, the angles
81 and 82 being equal.
2. All a lines (f31ines) turn through the same angle in going from one~ line
(a line) to another.
3. If one a line (f3 line) is straight between two ~ lines (a lines), then all
a lines (f3lines) are straight between these two~ lines (a lines). Furthermore,
these straight segments have the same length.
4. If both the a and ~ lines are straight in a certain region, all the stresses
in the region are constant. This is called a field of uniform stress state.
5. Along a straight shear line the state of stress is constant.
6. If the state of stress is constant along a curve, then either the curve is
embedded in a field of constant stress or else the curve is a straight shear line.
7. The radii of curvature of the a lines (f3lines) where they intersect a given
~line (a line) decrease in direct proportion to the distance traveled in the
positive direction of the f3line (a line). Therefore, if the plastic zone extends
far enough, the radii of curvature eventually become zero, so that neighboring
slip lines run together and the solution ends at the envelope of the slip lines.
This is Hencky's second theorem.
8. As we proceed along a given slip line of one family, the centers of curva-
ture of the slip lines of the other family form an involute of this slip line.
9. The envelope of the slip lines of one family is the locus of the cusps
of the slip lines of the other family.
270 The SUp-Line Field [Ch. 12
10. The envelope of the slip lines of one family is a limiting line across
which the shear lines of the other family cannot be continued.
11. If the radius of curvature of an a line (ftline) jumps discontinuously as
it crosses a {3line (a line), all a lines ({3lines) crossing the {3line (a line) will
suffer the same jump in radius of curvature. This also means that the deriva-
tives ofthe stresses are discontinuous across the slip line.
There are many similar theorems, but they are not of practical interest.
Hencky's first theorem can easily be proved as follows. Referring to Figure
12.3.1, along the a line AD, the first of Hencky's equations (12.1.18) gives
Therefore,
(12.3.1)
Also along AB,
xA + eA = xB + eB
and along BC
Xa - Ba = XB - eB
Therefore,
(12.3.2)
(12.3.3)
Theorem 3 is also a direct corollary, since if one of the lines is straight, say
AD, then Bn - ()A = 0 and therefore 80 - (JB = 0, so the other line is also
straight.
Similarly Theorems 4 through 6 follow directly from Hencky's equations
(12.1.18). Theorems 7 through 10 are based on the theory of plane curves.
Thus let the radii of curvature of the a and {3 lines at the point A in Figure
12.3.2 be designated by Ra and Rp, respectively. At the point B the
Sec. 12-3] Geometry of the Slip-Line Field
Rrt.-6f3
Rrt. (3
--r~
~----~r-~-----
8 c
0 ----- 1
1
of curvature R" has decreased by an amount Af3, and at the point D, Rp has
decreased by an amount Ao:, to first order of small quantities. In the limit 1
1
(12.3.5)
1
8Rp = -l
00:
1
or, more conveniently for computational purposes, since Ao: = Ro:Ae and
Af3 = -RpAO, equations (12.3.5) can be written 1
neighboring slip lines becomes zero, and so does the radius of curvature of
the (3 line through T, as stated in Theorem 7. It is a point on the envelope of
the a lines and is a cusp of the (3 line.
If the stress is constant throughout the field, the slip lines form two sets
of orthogonal straight lines. This follows directly from Hencky's equations
(12.1.18), for if X is constant, then 8 is constant. This is the converse of
Theorem 4 of Section 12.3.
Centered Fan
((lines
the straight lines be the a lines and the circular arcs the f3 lines. Then from
Hencky's first equation, since 8 is constant along an a line, X must also be
constant along an a line, and, from the second equation, since 8 varies linearly
with distance along a f3 line, X must vary linearly with distance along a f3line.
Thus the mean stress is constant in the radial direction and varies linearly
with the angle measured from the x axis. To find the stress components we
then make use of equations (12.1.14). This type of slip-line field is called a
centered fan. Note that the center of the fan is a singular point of the stress
field, since it can have any one of an infinity of values.
Indentation by a Punch
Y ---C( lines
----13 lines
' '\
'\
F '
and body are perfectly lubricated, so that there is no friction between them.
It is also assumed that there is a constant pressure over the face of the punch.
For the boundary conditions we have that over the segment AB there is a
uniform pressure kp, and the rest of the boundary is stress-free. We consider
only the case of incipient plastic flow, since once plastic flow progresses, the
shape of the boundary GABC changes considerably and it is necessary to
satisfy the boundary conditions on the deformed boundary.
Assume now that plastic flow occurs over a segment AG of the free bound-
ary as shown. The length of this segment is as yet not known. From the
boundary condition on this segment
onAG
274 The Slip-Line Field [Ch. 12
it follows that
ax= ±2k
onAG (12.4.1)
Since the shear stress is zero, AG is a principal direction and the slip lines
must be at ± 45° with AG. This also follows from the last of equations
(12.1.14). The a lines make 45° angles with AG and the {31ines 135° (or -45°),
as shown in Figure 12.4.2.
Consider the triangular region AGFformed by AG and the slip lines AFand
GF. By Theorem 6 of Section 12.3, this is a constant stress regipn. The slip
lines are straight lines with 8 = 71"/4. The mean stress X is a constant and must
satisfy equations (12.1.18) throughout this region. Since on the boundary AG
it follows that
X- _1.} 2
8 =i inAGF (12.4.2)
Now consider the boundary AB. Since it has been assumed that there is
no friction, r = 0 along this boundary, so that
a =
: =
-kp} alongAB
0
(12.4.3)
Therefore,
(ax~ ayr = k2
or ax= ay ± 2k = k(2- p)
where the plus sign has been chosen (see Problem 7). It follows then, just as
for the segment AG, that AB is a principal direction and the slip lines make
± 45° angles with AB. This time the {3lines are at 45° with AB and the a lines
Sec. 12-4] Some Simple Examples 275
are at 135° (or -45°). As before, region AEB is a constant stress region, the
slip lines being straight lines with 8 = ;br. From the boundary condition,
X=
Ux + Uy
4k = -2-
1- p} inAEB (12.4.5)
8= tn-
Now AF and AE are straight a slip lines and it follows from Theorem 3 of
Section 12.3 that all the shear lines in between these two are straight, or
region F AE is a centered fan. The stresses are then constant along any radial
line from A to the arc FE and vary linearly along any arc such as IJ from the
value
1-p
X= -! along AF to X= - -
2
along AE
X +8= constant
Along HI,
X=-! and
AlongJK,
X= !(1 - p) and
Hence
-! + i = !(1 - p) + tn-
or p=2+7T (12.4.6)
In the above example of the punch indentation, the slip lines and the
solution were obtained completely in closed form from the boundary condi-
tions. In general, however, numerical or graphical methods will be necessary.
In this section a brief discussion of the simplest numerical methods will be
presented. For this purpose we must distinguish among three types of
boundaries, as shown in Figures I2.5.I through I2.5.3. Figure 12.5.1 shows
a case where the boundary curve C0 is not a slip line. The values of x and 8
are given on this boundary and it is desired to construct the slip-line field~
We choose a number of stations on the arc C0 and try to construct the
lines passing through these points. The various shear lines are uvo•~J<.~'"
1, 2, 3, etc., and a grid point at the intersection of the ith a line with the jth
{3 line is designated i,j.
Now consider the points (I, I) and (2, 2) on the curve C0 • At the
(1, 1) draw a straight line with angle 811 , representing the a line through
point. At (2, 2) draw a line with angle 822 + n/2. The two lines intersect
Sec. 12-5] Numerical Solutions of Boundary-Value Problems 277
y
Co
L-----------------------------------~x
FIGURE 12.5.1 Numerical solution of first boundary-value problem.
(1, 2), which is a rough approximation to the true intersection point of the
a and f3 lines. (If the a and f3 lines happened to be straight lines, the inter-
section point would be exact.) The points (2, 3), (3, 4), (4, 5), etc., can be
determined approximately the same way. From Hencky's equations we have
Xu - Ou = X12 - 812
(12.5.1)
X22 + 822 = X12 + 812
(12.5.2)
We can now proceed to find X and 0 at (1, 3), (2, 4), (3, 5), and (4, 6). We thus
obtain the slip-line field and the stresses in the entire region bounded by C0
and the terminal slip lines AP and QP. A little reflection indicates that the
solution cannot be carried beyond region APQ without some additional
information. This leads us to the following theorem. Given an arc C0 which
is not a slip line and all the stresses acting at every point along the arc, then
the complete slip-line field and the corresponding stresses can be determined
278 The Slip-Line Field [Ch. 12
within the region bounded by C0 and the intersecting terminal slip lines as
shown in Figure 12.5.1. Region APQ is called the region of influence of the
arc Co.
If the arc C0 is itself a slip line, then the previous method is obviously
inapplicable. If a second slip line, intersecting the first one, as shown in
Figure 12.5.2, is also given, a solution can be obtained. For if 8 is known
c
L---------------------------~~x
FIGURE 12.5.2 Second boundary-value problem.
along both slip lines, then 8 can be determined at the adjoining net points by
use of Hencky's first theorem. The complete slip-line field can then be con-
structed within the quadrilateral shown in Figure 12.5.2. To determine the
stresses, it is necessary to know the value of X at just one point on the bound-
ary slip line, for by use of Hencky's equations (12.1.18) X can then be
computed throughout the region.
Alternatively, if the curve C0 is a slip line, a solution can be obtained if on a
second intersecting curve, not a slip line, either X or 8 is specified. The solution
can then be obtained in the region indicated in Figure 12.5.3, using techniques
similar to those previously outlined. Details of the solution techniques for all
three types of boundary-value problems, as well as methods for improving
the accuracy, can be found in references [1], [4], and [5].
A geometric construction for the stress and velocity fields, which is fre-
quently very useful and leads to a better insight into the principles underlying
slip-line theory, has been suggested by Prager [6]. For this purpose we make
use of two planes, called the stress plane and the physical plane, as shown in
Figure 12.6.1. Consider a point P undergoing plastic flow. The stress vector
y
II
~---rrx
(a)
TI (b)
acting at the point P will depend on the orientation of the area element
through the point P upon which it acts. This is shown in the physical plane
of Figure 12.6.1(b). The figure shows the traces of several area elements whose
~normals lie in the xy plane. These area elements actually contain the point P
but are here shown separated for clarity. The shaded side of a given trace
represents material, and the stresses shown are those transmitted from the
tlnshaded side to the shaded side. Instead of identifying an area element by
.he direction of its normal, it is convenient to identify it by the direction of
280 The Slip-Line Field [Ch. 12
the trace of the element on the xy plane. Thus angles will be measured
counterclockwise from the negative y axis instead of the x axis.
On the stress plane, Figure 12.6.1(a), a Mohr circle is plotted for the
stress state at the point P of the physical plane. The Mohr's circle is con-
structed using the following convention. A shear stress which will cause the
element to rotate in a clockwise sense is considered as positive, counter-
clockwise as negative. Thus the stress state (ax, - T xy) on plane P A is shown
as point A of the stress plane. On plane PB, whose normal stress is ay, the
shear stress is positive (clockwise) and (ay, Txy) gives the point Bin the stress
plane. We note that the angle</> between the resultant stress Sand the normal
stress Sn in the stress plane is the same as the angle between the resultant
stress and the normal stress in the physical plane, but it is measured in the
opposite direction; i.e., if <{> is measured clockwise in the stress plane it is
measured counterclockwise in the physical plane as shown.
Once the points A and B are determined, the Mohr's circle can be con-
structed in the usual way by drawing a circle through A and B whose center
lies at the intersection of AB with the Sn axis. The stress vector acting on any
plane through P can now be found from the following consideration. If a
straight line is drawn in the stress plane through any stress point (e.g., A or
B) parallel to the trace in the physical plane upon which the stress acts, it
will intersect the Mohr's circle at a point P, and all such lines will intersect
the circle at this same point. The point P is called the pole of the Mohr's
circle. In Figure 12.6.1(a) the pole is obtained by drawing a vertical line
through A to the point P or a horizontal line through B. For example, to
obtain the stress vector on plane OC rotated a degrees counterclockwise from
OA, one draws a line through P on the Mohr's circle parallel to OC or a
degrees counterclockwise from P A. The point of intersection of this line with
the Mohr's circle gives the state of stress on the plane PC.
The top and bottom points of the Mohr's circle, labeled I and II, corre-
spond to the planes upon which the maximum and minimum shearing stresses
± k act. The directions of the traces of these planes are given by the
PI and PII. These directions are the first and secop.d shear directions or
a and (3 directions, respectively, as defined in previous sections.
The geometric construction of the slip-line field can now be o
making use of the following fact, proved by Prager in reference [6]. As
move along a slip line in the physical plane, the pole of the Mohr's
traces out a cycloid in the stress plane. It does this by rolling without
along the top tangent T = k if we move along an a line and along the
tangent T = - k for a (3 line. Thus assume the stress state is known at
point Pin the physical plane of Figure 12.6.2(b). The Mohr's circle and
pole P can be constructed as shown in Figure 12.6.2(a). The directions
Sec. 12-6] Geometric Construction of Slip-Line Fields 281
I I I
1st
'\ P'
\
\
(a) (b)
FIGURE 12.6.2 Cycloid trace of pole on stress plane (a) and corresponding slip
line in physical plane (b). (Reference [5].)
the first and second slip lines at Pare also known from PI and PII. If we now
move along the second shear line (13 line) to the point P', the pole P of the
Mohr's circle will move toP' as the circle rolls on the tangent lines. = -k.
The tangents to the two slip lines atP' are given by the directions P'IandP'II,
as shown by the dashed lines in Figure 12.6.2(a). Alternatively, since
P'II is normal to the cycloid at P' (the point II is the instantaneous center of
rotation), the element of the slip line at P' in the physical plane is normal to
the element of the cycloid at P' in the stress plane. The slip lines at the point P"
can be established the same way. At the same time the stresses a~ and a~ are
determined from the positions of the center of the circle. It is apparent from
the above that cycloids generated in this fashion are the images in the stress
plane of the slip lines in the physical plane.
Prager has also shown [6] how to construct the velocity field at the same
01
(a) Slip lines ( b) Hodograph
time as the slip lines. Let P 1 , P 2 , and P 3 be neighboring points on the first
slip line, as shown in Figure 12.6.3. The line P 1 P 2 is in the first shear direction.
Let Va and v13 be the velocity components in the two slip directions at P 1 and
let Va + dva and v13 + dv 13 be the corresponding velocities atP2 • The condition
that there be no extensions, only shearing flows in the slip directions [see
equation (12.2.4)], requires that the projections of the velocity vectors at the
points P 1 and P 2 onto the line P 1 P 2 must be equal. This yields immediately
the Geiringer equation (12.2.8):
( 0) (b)
possible to draw two circles for each of the points. For example, in Figure
12.6.5 the stress vectorS acting at a point P has been drawn from the origin
to the point A of the stress plane. Through the point A, however, two circles
can be drawn, each with radius k and with their centers on the Sn axis, as
shown. One of these corresponds to a lower mean normal stress than the
other and these solutions are therefore called the weak and strong solutions,
respectively. As indicated on the figure, the poles and the shear directions will
be different for the two circles. Which of these solutions to choose is usually
determined from additional boundary conditions or from looking at the
boundary conditions as a whole, rather than at one point. This uncertainty
is due to the quadratic nature of the yield criterion, as was indicated for the
punch indentation problem of Section 12.5.
Returning now to Figure 12.6.4, the weak solution is adopted. The Mohr's
circles for points A and B are shown by the solid and broken-line circles,
respectively. As the solid-line circle rolls along the line r = k, the pole will
trace a cycloid which is the image of the first slip line through A. As the
broken-line circle rolls along s. = - k, the point B will describe a cycloid,
shown by the broken line, which will be the image of the second slip line
through B. The intersection of the two cycloids at C will be the image of the
intersection of the two slip lines in the physical plane at C. The directions of
the normals to the cycloids at C fix the directions of the tangents to the slip
lines at C in the physical plane. The actual location of C in the physical plane
is found from the conditions that AD = CD and BE = CE. This amounts
to assuming that the slip-line arcs AC and BC are circular.
Once the slip lines have been established, the hodograph can be constructed
from the known velocities on the boundary using the orthogonality relation
between the slip line and hodograph elements.
The above is a brief outline of Prager's geometric construction of slip-line
fields. The interested reader is referred to references [6] through [9] for further
discussion and examples.
to determine if the stress fields in the rigid regions also satisfy equilibrium
and the boundary conditions and nowhere exceed the yield stress. If such a
statically admissible stress field exists in both the plastic and rigid parts, then
this solution constitutes a lower bound, and since the solution is both an
upper bound and a lower bound, it is a complete solution, giving the true
load for the problem.
The fact that a statically admissible stress field as defined above gives a
lower bound is expressed by the lower bound theorem [10], which can be
stated as follows: A load which produces a statically admissible stress field will
be equal to or less than the true load that will produce plastic flow.
Lower bound solutions are usually difficult to obtain since it would be
necessary to solve the equilibrium equations in the elastic or rigid region.
Upper bound solutions can be obtained from slip-line solutions and are of
greater value, in as much as they ensure that a certain operation will be
performed, since the calculated load will be greater than the required load.
Discussions of methods of obtaining upper bounds with examples can be
found in references [II] and [5].
The slip lines previously discussed are actually the characteristics of the
differential equations defining the problem. All the properties of charac-
teristics of hyperbolic equations and all the mathematical methods for
solving such equations can therefore be applied directly to the plane strain
problem of a rigid-perfectly plastic material. In this section we shall briefly
describe the origin and properties of characteristics to give the reader an
insight into the mathematical origin of slip-line fields. The reader interested
only in the physical description of the previous sections may skip to Chapter
13 without loss of continuity.
Consider the first-order differential equation
8u 8u
a-+b-=c (12.8.1)
ex ey
where a, b, and c are functions of u, x, andy but not of the partial derivatives
of u. Such an equation is called quasilinear. Introduce the standard notation
8u
8y =q (12.8.2)
286 The Slip-Line Field [Ch. 12
(12.8.4)
du 8u dx 8u dy dx dy
ds = 8x ds + 8y ds = p ds + q ds (12.8.5)
Equations (12.8.5) and (12.8.3) give us two equations in the two unknowns
p and q. Solving these two equations gives
cdy- bdu
p =a dy- b dx
adu- cdx
q =a dy- b dx
Thus, given a, b, and c, the value of u on the curve C and the shape of the
curve, p and q can be computed from (12.8.6) and u in the neighborhood
of C obtained from (12.8.4). However, this will not be true for any curve C,
for if Cis such that the denominator in (12.8.6) vanishes; i.e., if
dy b
dx =a
Sec. 12-8] Slip Lines as Characteristics 287
then there will obviously be no solution unless the numerators also vanish.
A curve C whose equation satisfies (12.8. 7) is called a characteristic curve,
and if u is specified along such a curve, there will be no solution to the
problem unless u is specified so that
du c
or
dx =a (12.8.8)
(12.8.10)
or
u= x +B = y + (B - A) (12.8.11)
along these lines. Suppose now that u is given on the line segment 0 < x < 1,
y = 0. This line segment is the curve C of the previous discussion. Then
since Cis not a characteristic, we should be able to obtain a unique solution
for u. The simplest way to do this is to integrate along the characteristics.
Th11s considering the characteristic intersecting the x axis at x = x1 as shown,
288 The Slip-Line Field [Ch. 12
the value of u specified at this point is U;, and since along this characteristic u
must satisfy (12.8.11 ), it follows that
B-A=Ut
so that
U = Y + Ut = X + Ut - Xt (12.8.11a)
for all
0 <X< 1
A solution has thus been obtained for the problem using the initial data
and the characteristic and compatibility equations (12.8.10) and (12.8.11).
These equations are the equivalent of the original differential equation,
which does not have to be used at all. This method of solving a partial differ-
ential equation by reducing it to a set of ordinary differential equations for
finding the characteristics and for integrating along the characteristics is
called the method of characteristics. Thus the introduction of characteristics
not only serves the purpose of determining whether a unique solution exists
but actually provides a method for obtaining the solution.
From the above example it is apparent that the solution is defined and is
unique only in the region bounded by the terminal characteristics at x = 0 and
x = 1. This region is called the region of influence. To determine the solution
outside this region requires additional data outside the range 0 < x < 1.
We can state this result as a general theorem.
A partial differential equation having real characteristics has a unique solution
within the region bounded by the curve C upon which the initial data is specified,
and the two characteristics intersecting the ends of C. If C coincides with
characteristic there is no unique solution.
Let us consider now the case when u is specified along one of the
teristics, say, the line x = y. For
specified on this line such that
U = X + Uo = Y + Uo
Sec. 12-8] Slip Lines as Characteristics 289
will satisfy the differential equation (12.8.9) and also the initial values on
y = x. This illustrates the fact that if the initial conditions are specified on
characteristic, the solution is not unique. In this case we might say that the
solution is not defined uniquely at points not on this line because the terminal
characteristics coincide, and the region of influence is just the line itself.
Another very important property of characteristics can now be seen-the
possibilities of discontinuities in the solution which are propagated along
characteristics. For example, suppose on the initial line y = 0 of the previous
example we are given
u = 1
u=x+1
8u =0 8u = 1 to the left of y = x - t
ox 8y
(12.8.16)
8u = 1 Bu =0 to the right of y = x - t
Bx oy
290 The Slip-Line Field [Ch. 12
Second-Order Equation
(12.8.17)
(12.8.18)
Ar + Bs + Ct = D (12.8.19)
Now assume that on some initial curve C0 in the xy plane, the values of u,
and its normal derivative are given. Alternatively the partial derivatives p and
q are given, for then u and the normal derivative can be computed. The
values of u, p, and q along C0 are called a strip of first order and u, p, and
must obviously satisfy the condition
du = p dx + qdy
called the strip condition.
We now ask ourselves the same question as before. Does there exist
solution u of equation (12.8.17) satisfying these initial conditions? If
second derivatives r, s, and t on the curve can be determined so as to
Sec. 12-8] Slip Lines as Characteristics 291
du = -dx
au + -dy
au
ox ay
au) 8 u 8 u 2
d (oy = ox oy dx + oy2 dy
2
or rdx + sdy = dp
sdx + t dy = dq
(12.8.20)
and the differential equation is
rA + sB + tC = D
A, B, C, D, dq, and dp, being known on the curve, we have three equations
to solve for the unknown second derivatives r, s, and t. For example,
dx dp 0
0 dq dy
A D c
S= (12.8.21)
dx dy 0
0 dx dy
A B c
If the determinant in the denominator vanishes there will be no solution
unless the numerator also vanishes, in which case there are an infinite number
of solutions. If the denominator vanishes,
A(dy)2 - B dx dy + C(dx) 2 = 0
(12.8.21a)
(dy)
2
dy
A- -B-+C=O
dx dx
dy B ± VB 2 - 4AC
(12.8.22)
dx = 2A
292 The Slip-Line Field [Ch. 12
Equation (12.8.22) defines two sets of curves (one using the plus sign, one
using the minus sign). These are called characteristic curves, and if the data
are given along one of these curves, no solution exists unless, as already
mentioned, the numerator determinants also vanish. For this case,
A dp dy + C dq dx - D dx dy = 0 (12.8.23)
In the hyperbolic (and parabolic) case, if the curve Cis not a characteristic,
a solution exists. The prescribed values in this case need not be regular and
the higher derivatives of the prescribed values may have finite jumps at points
on C. These discontinuities are then propagated along the characteristics
where they originate. If Cis a characteristic curve, there is no solution unless
u and its derivatives satisfy a compatibility relation, in which case there are
an infinity of solutions. It is also evident that closed boundaries are excluded,
for a given characteristic would then intersect the boundary twice, and values
of u and its derivatives could not be assigned arbitrarily to both points of
intersection.
Boundary-Value Problems
The reader will recognize the similarity between these three boundary-value
problems and the three boundary-value problems for slip-line fields discussed
in Section 12.5.
Finally, it follows that by combining these three basic types of boundary
conditions, more complicated conditions can be used. For example, in
Figure 12.8.2(d), u, p, and q may be given on C0 • But then only one relation
of the type a(8uj8n) + bu = c(x, y) can be prescribed on each of Cb and C~.
294 The Slip-Line Field [Ch. 12
x=constant x=constant
jl=constont
(a) (b)
jl=constont
(c) (d)
The solutions in the various regions shown can then be patched together along
characteristics so that u is continuous, while 8uj8x and 8uj8y may be dis-
continuous. Note again that closed boundaries are not admissible.
Numerical Solution
simple problems have been solved in this way. The widest use of the method
of characteristic is found in conjunction with numerical methods.
Suppose then that u, p, and q are given along a noncharacteristic arc C and
we wish to obtain the solution in the region bounded by C and the two
terminal characteristics as shown in Figure 12.8.2(a). It is desired to solve
equations (12.8.22) and (12.8.23) numerically. If A, B, and C are functions
only of x and y, (12.8.22) can be solved separately and the characteristics
constructed independently of the initial conditions. Such characteristics are
called fixed. If A, B, and C are also functions of u, p, and q, then (12.8.22)
and (12.8.23) must be solved simultaneously. (Note: If A, B, and C are
functions of u, p, and q but not of x andy, then it is possible to interchange
the roles of the independent and dependent variables and transform the
partial differential equation into a linear differential equation with x and y
as the dependent variables. This is called a hodograph transformation. The
solution is then obtained in the plane of u and 8u/8n, or p and q, called the
hodograph plane, instead of the xy plane.)
Equations (12.8.22) can be written
dy+ =f+ dx
(12.8.24)
dy- =f- dx
where
B + VB 2 - 4AC
2A
(12.8.25)
Let us then draw the curve C and two families of characteristics as shown
in Figure 12.8.3. Consider any two adjacent points P and Q on C and let the
j+ characteristic from P intersect the f- characteristic from Q at R. If the
characteristics are fixed, the position of the point R is known. If the charac-
teristics depend on the solution, the coordinates of R must be determined at
the same time that p and q are determined at this point. Equations (12.8.24)
and (12.8.26) are now written approximately as follows:
Equations (12.8.27) and one of (12.8.28) furnish five equations for the five
unknowns xR, yR, PR, qR, and uR. These equations are nonlinear and they
usually have to be solved by some iterative method. Convergence is usually
rapid if the interval PQ is not too large.
In the same way, the solution at other grid points adjacent to the initial
curve, such as Sin the figure, can be obtained. From R and S we can proceed
to T, and so on. The solution is of course defined only in the region bounded
by the terminal characteristics, as previously explained.
au au & &
a 8x + b By + e 8x + f By =d
A 8u + B 8u + E dv. + F 8v = D
8x By 8x By
Sec. 12-8] Slip Lines as Characteristics 297
with
au au
du = ox dx + dy dy
(12.8.30)
ov ov
dv = ox dx + oy dy
then by a similar procedure the equation for the characteristics is
2
dy) dy .
(aE - eA) ( dx + (eB +fA - bE - aF) dx + bF - fB = 0 (12.8.31)
du dx dy 0
dv 0 0 dx
=0 (12.8.32)
d a b e
D A B E
It will now be shown that the slip lines discussed in Sections 12.1 through
12.7 are characteristics of the governing differential equations and all the
properties of characteristics discussed in Section 12.8 apply to them. We
start with equations (12.1.16), which are essentially the equilibrium equations
with the yield condition included:
ox 88 . 88
-
ox
- COS 28- -
dx
Sill 28-
oy
=0
(12.8.33)
ox . 88 88
-
ay
- Sill 28 -
ax
+ COS 28 -
ay
=0
These are two simultaneous quasilinear first-order equations, and if they are
hyperbolic we should be able to use the method of characteristics. Comparing
the coefficients with those in (12.8.29), it is seen that
sin 28 (;7x)
2
+ 2 cos 28 1x - sin 28 = 0
dy -cos 28 ± 1
or (12.8.34)
dx = sin28
298 The Slip-Line Field [Ch. 12
or
2 sin 2
. a
ea= tan
l)
u
2 Sin u COS u
(12.8.35)
But tan 8 and tan(& + 'TI'/2) are the slopes of the slip lines, so that (12.8.35)
shows that the characteristics of the differential equations (12.8.33) are just the
slip lines. We can identify the ex slip lines with the plus characteristics and the
(3 slip lines with the minus characteristics, as defined by equations (12.8.35).
To determine the compatibility relation that must be satisfied along the
slip lines, equation (12.8.32) gives
dx dx dy 0
de 0 0 dx
=0 (12.8.36)
0 0 -cos 28
0 0 -sin 28
Problems
1. Derive equation (12.1.5).
2. Show that the principal stresses for plane strain problems are given by
equations (12.1.7) and (12.1.8).
3. Show that the Tresca criterion differs from equation (12.1.5) only by a
multiplicative constant.
4. Derive equation (12.1.9).
5. Show that equations (12.1.14) can be obtained by means of equations
(12.1.4) and (12.1.12).
6. Prove Theorems 6 and 8 of Section 12.3.
7. Determine the velocity distribution for the condition of Figure 12.4.2 using
the Geiringer equations.
8. Obtain the complete solution for the rigid punch indentation of Figure
12.4.2 assuming the rigid plastic boundary to be HIJKLMN.
Sec. 12-8] Slip Lines as Characteristics 299
9. Show that the choice of a plus sign in equation (12.4.1) or a minus sign in
equation (12.4.4) would lead to solutions which violate the initial assump-
tions.
10. Prove that the point of intersection of a line drawn through the pole of the
Mohr's circle and the circle will give the stress state acting on a plane whose
trace is parallel to this line.
References
1. W. Prager and P. G. Hodge, Jr., Theory of Perfectly Plastic Solids, Wiley
New York, 1951.
2. P. G. Hodge, Jr., An Introduction to the Mathematical Theory of Perfectly
Plastic Solids, ONR-358, Feb. 1950.
3. L. Prandtl, Dber die Harte Plastischer Koerper, Goettinger Nachr., Math.
Phys. Kl., 1920, pp. 74~85.
4. R. Hill, The Plastic Yielding of Notched Bars Under Tension, Quart. J.
Mech. Appl. Math., 2, 1949, pp. 40-52.
5. E. G. Thomsen, C. T. Yang, and S. Kobayashi, Mechanics of Plastic Defor-
mation in Metal Processing, Macmillan, New York, 1965.
6. W. Prager, A Geometrical Discussion of the Slip-Line Field in Plane Plastic
Flow, Trans. Roy. Inst. Tech. (Stockholm), 65, 1953, pp. 1~26.
7. W. Prager, The Theory of Plasticity: A Survey of Recent Achievements,
Proc. Inst. Mech. Eng., 169, 1955, pp. 41~57.
8. H. Ford, The Theory of Plasticity in Relation to Engineering Application,
J. Appl. Math. Phys., 5, 1954, pp. 1~35.
9. J. M. Alexander, Deformation Modes in Metal Forming Processes, Proceed-
ings of the Conference on Technical Engineering Manufacture, Institute of
Mechanical Engineers, New York, 1958, Paper No. 42.
10. D. C. Drucker, H. J. Greenberg, and W. Prager, The Safety Factor of an
Elastic~Plastic Body in Plane Strain, J. Appl. Mech., 18, 1951, pp. 371~378.
11. W. Johnson and P. B. Mellor, Plasticity for Mechanical Engineers, Van
Nostrand, Princeton, N.J., 1962.
12. L. Fox, Numerical Solution of Ordinary and Partial Differential Equations,
Pergamon Press, London, 1962.
13. G. A. Korn and T. M. Korn, Mathematical Handbook for Scientists and
Engineers, McGraw-Hill, New York, 1961.
14. H. Hencky, Ueber einige statisch bestimmte Faelle des Gleichgewichts in
plastischen Koerpern Z. angew. Math Mech., 3, 1923, pp. 245~251.
15. H. Geiringer, Beit zum Vollstandigen ebenen Plastizitats-problem, Proc 3rd
Intern. Congr. Appl. Mech, 2, 1930, pp. 185~190.
General References
Prager, W., Introduction to Plasticity, Addison-Wesley, Reading, Mass., 1959.
Ford, H., Advanced Mechanics of Materials, Wiley, New York, 1963.
CHAPTER.13
LIMIT
ANALYSIS
The theory of limit analysis is used primarily in the design of steel structures
composed of various elements such as beams, frames, girders, arches, etc.
For many years the basis for structural design has been the allowable stress
concept. The allowable stress was usually taken to be the yield stress of the
material, and the design stress was then taken to be some fraction of the
allowable stress, depending on the factor of safety used. In some applications
the allowable stress was governed by the possibility of buckling or fatigue.
The design methods used were always elastic.
It is apparent, however, that the important consideration in an engineering
structure is not whether the yield stress is exceeded at some point, but whether
the structure will carry the intended loads or perform its intended function,
and there is really no reason for assuming that the stress in the structure
should never exceed the elastic limit. As a matter of fact, it is fairly evident
that in almost all structures, local plastic flow will occur at stress raisers and
at points of discontinuity in the geometry, and, furthermore, residual stresses
as high as half the yield strength may already be in some elements as they
come from the steel mills, before the load is even applied.
The practice is therefore becoming more widespread to design structures
into the plastic range, the materials being assumed to behave in an elastic-
perfectly plastic manner. In these design procedures no attempt i~ made to
300
Sec. 13-2] Simple Truss 301
determine the stresses and strains in the structure, but rather what is sought
is the load-carrying capacity or limiting load at which the structure will
collapse. This type of analysis is called limit design or plastic design, and the
load at collapse is called the plastic collapse load.
For the great majority of problems this type of approach makes more
sense than a design based on an elastic analysis. Furthermore, since the stress
distribution is not SOJ.lght, it is much simpler. There are other important
problems, however, which cannot be resolved on the basis of simple limit
analysis. Among these are buckling, fatigue, and fracture.
The first use of plastic design in structures was apparently made by
Kazinczy in the design of apartment buildings in 1914 (see reference [1]).
Since then many contributions have been made to this theory both in this
country and abroad. Plastic design is already part of certain specifications in
some countries, and it is being used more and more by engineers in this
country.
2 2
L
______ lI
T2
F
FIGURE 13.2.1 Simple truss.
l13.2.1)
302 Limit Analysis [Ch. 13
(13.2.2)
(13.2.3)
(13.2.4)
which results in
(13.2.5)
Substituting into (13.2.1) gives
F
T2 = 2 + V2
(13.2.6)
2F
Tl = 2 + V2
T1 = 11,720 lb
(13.2.7)
T2 = 5,860 lb
Sec. 13-2] Simple Truss 303
This completes the elastic analysis, the stresses and deformation having been
determined. If the question is now asked, How strong is the truss?, the usual
answer is to quote a factor of safety giving the ratio of the maximum safe
load to the applied load. The conventional practice has been to take the
maximum safe load as that load which would just cause the maximum stress
to reach the yield stress. In this case, therefore, the factor of safety would be
a0 30,000
f = T 1 /A = 11,720 = 2 · 56 (13.2.9)
This is the plastic collapse load. The deflections just as this load is reached
will be
o2 = 0.017 in.
(13.2.12)
o1 = v'.2 o2 = 0.024 in.
304 Limit Analysis [Ch. 13
which are still small. If the load is increased beyond 72,400 lb, large deforma-
tions will take place and it can be assumed that the structure is no longer
usable.
We note that the factor of safety based on the load-carrying capacity is,
for the case F = 20,000,
72,400
fp = 20 000 = 3·62
'
rather than 2.56. Moreover, the load-carrying capacity or the plastic collapse
load is determined simply from the equilibrium equation (13.2.1) and a
knowledge of the yield stress. The detailed elastic analysis previously made
was not necessary. This is another great advantage of limit analysis.
Now consider the inverse problem, i.e., the problem of design rather than
analysis. Supposing we were given a truss such as shown in Figure 13.2.1,
and told to design such a truss to have minimum cross-sectional areas for a
working load of 20,000 lb, and a factor of safety of 3, taking all three bars
to have equal areas.
According to elastic design, the forces in the bars at a load of 3 x 20,000 =
60,000 lb would be, from (13.2.6),
60 000
T.2 = •
3.414
= 17 600 lb
'
T1 = 35,200 lb
To keep the stress in the middle bar below the yield stress of 30,000 psi, the
area would have to be
A 35,200 1 17 . 2
= 30,000 = . lll.
If plastic analysis were used, however, the stresses in all three bars could
be allowed to reach 30,000 psi at a load of 60,000 lb. Therefore, from
(13.2.1),
or A .i
= 2 14 = 0.83 in. 2
Sec. 13-3] Pure Bending of Beams 30_5
Thus, using plastic design, 30 per cent of the material is saved. Furthermore,
under the working load of 20,000 lb, the actual stresses in the bars are
al = Tl = 11,720 = 14 100
A 0.83 '
a2 = 2= 7,050
The next problem considered is the pure bending of a beam. Consider the
beam shown in Figure 13.3.1. The conventions for positive moments and
With the usual assumption of beam theory that plane sections remain plane
the strain ex is given by
Bx = kz (13.3.2)
(13.3.3)
(13.3.4)
306 Limit Analysis [Ch. 13
The maximum stress will occur at the outer fibers and. will be tensile at
z = h and compressive at z = -h. When this maximum stress equals the
yield stress a 0 , we have, from (13.3.3),
ao = kEh (13.3.5)
and substituting into (13.3.4) gives for the moment when the outer fiber just
reaches the yield stress
(13.3.6)
Me, (b) shows the distribution when M just equals Me, and as M is increased
more of the beam reaches the yield stress but the part between rh and - rh
is still elastic, as shown in (c). The moment for the condition shown in (c) is,
from (13.3.1),
rh kEz 2 dz + 4b
fh
M = 2b
J
-rh r/1
a0 z dz
but, from (13.3.5),
kE = ao
rh
Hence
Sec. 13-4] Beams and Frames with Concentrated Loads 307
Comparing equations (13.3.8) and (13.3.6), it is seen that the ratio of the
fully plastic moment to the maximum elastic moment is
~: = 1.5 (13.3.9)
This ratio is called the shape factor of the beam. It will, of course, be differ-
ent for different-shaped cross sections. For a circular cross section, for
example, it will be 1. 70. Figure 13.3.3 shows the values for several common
beam sections.
In Section 13.3 the case of pure bending where all sections of the beam
behaved the same way was discussed. We now consider concentrated loads
where the moment distribution varies along the beam. Our purpose, as before,
will be to determine the collapse loads of the structure.
308 Limit Analysis [Ch. 13
~A fF
L~----)o')o-1
/ <
(a)
M = LFo = Mo (13.4.1)
the section at A will become fully plastic. Since perfect plasticity has been
assumed, the beam will now be able to rotate freely about the point A and
will thus have collapsed. We say that a yield hinge or plastic hinge has formed
at A. Such a hinge can be thought of as one which is locked in place as long
as the moment M is less than M 0 , and which becomes free to turn freely when
M becomes equal to M 0 • The collapse load, from (13.4.1), is obviously
2
Fo = M 0 = 2bh a0 (13.4.2)
L L
for a rectangular cross-sectional beam.
Indeterminate Beam
F
L ---------;l>-1
A;~-<----L---)o-K---
'/!--------'--::--------;,
(a) B
(e)
the load F being applied at the middle. The elastic solution first requires
determining the reaction at C, which can be done by first calculating the
displacement at C with the support removed and calculating the reaction at
C necessary to bring the deflection back to zero with the force F removed.
The moment distribution can then be computed. To determine the collapse
load, however, none of this is necessary. The moment distribution when
everything is elastic is shown in (b), and as the force Fis increased, eventually
the section at A will become fully plastic, the moment at A being - M 0 • A
yield hinge is thus formed at A, just as in the case of the cantilever previously
discussed with the resultant moment distribution as shown in (c). Because
of the support at C, however, the beam will not yet collapse. As the load is
further increased, the point A cannot carry any more moment, but the rest
of the beam can and the moment at B will continue increasing until it reaches
the plastic moment M 0 , at which time a second yield hinge is formed. The
beam will then collapse, as shown in (e).
The calculation of the collapse load for this case is very simple. Let the
310 Limit Analysis [Ch. 13
Fa= 6bft2aa
L
Simple Frame
Ia) I b) I c) (d)
M4 = -2R2 L
If it is now assumed that the frame will collapse by the mechanism shown in
(c), then M 1 = -M0 , M 2 = M 0 , and M 4 = -Mo. From the second of
equations (13.4.5),
2L
F= Mo)
and from the first of (13.4.5), (13.4.6)
Ms = -!Mo
Ms = Mo
Then, from (13.4.5),
F= 2M0 (13.4.7)
L
312 Limit Analysis [Ch. 13
F= 5M0 (13.4.8)
8£
But this violates the original assumption for this case-that there is no yield
hinge at station 2; i.e., the moment at 2 is less than M 0 • This mechanism is
therefore also not admissible, leaving the mechanism of (c) as the only
possible one. The true collapse load is therefore given by equation (13.4.6).
Several results can now be noted. The collapse loads corresponding to the
cases of (d) and (b) are higher than the true collapse load given in (13.4.6).
It appears, then, that as the load is slowly increased from zero, the frame will
collapse the first chance it gets.
Second, if the load is decreased in these last two cases, so that the maximum
moment does not exceed M 0 , it would be necessary to divide by 12 and i,
respectively, as seen from (13.4.7) and (13.4.8), giving
and
FL 5
Mo 14
It follows then that as the load is slowly increased, the frame will not
collapse for any load for which some equilibrium configuration can be found,
unless it produces a mechanism.
These two concepts can be formalized into basic theorems of limit analysis
to be discussed in Section 13.5.
The previous problems were simple enough so that the correct collapse
load could be obtained in each case without too much difficulty. In more
complicated problems, however, it may be difficult or impossible to obtain
Sec. 13-5] Theorems of Limit Analysis 313
the collapse load exactly, and recourse must then be had to the upper and
lower bound theorems, which provide upper and lower bounds to the true
collapse load. These theorems are essentially the same as those presented in
Section 12.7 for the load to produce incipient plastic flow in a rigid-perfectly
plastic material. They were first presented by Gvozdev [3] and independently
proved by Hill [4, 5] for the rigid-perfectly plastic material, by Drucker et al.
[6, 7] for the elastic-perfectly plastic material, and in references (8] and [9] for
the special cases of beams and frames. We shall present the theorems in their
general form. The proofs, which are based primarily on the principle of
virtual work, can be found in the cited references and in a particularly elegant
form in reference [10].
may, or may not, increase the strength of a body, but it will certainly not
decrease it.
The upper and lower bound theorems can be stated in a more precise and
elegant fashion by first introducing the following definitions and concepts [2],
which are useful in their own right.
Generalized stresses Q1 are used to designate the state of stress in a body.
These may be actual stresses, so that Q 1 = ax, Q 2 = ay, etc.; or they may
be moments, as in the case of beams, so that Q1 = M; or resultant forces
and moments as in the case of shells, where Q1 = N 8 , Q2 = Nq,, Q3 = M 8 ,
etc. The choice of generalized stresses to be used depends on the particular
problem and is a matter of convenience and is not necessarily unique. Once
such a choice has been made, however, the corresponding generalized strains
q1 are defined so that the increment of internal work done is given by
(13.5. 1)
For example, for the beam problems considered, the generalized stress at a
hinge is M, and the corresponding generalized strain is the rotation 8, the
internal work at the hinge being equal to M8.
Instead of using the increment of internal work as given by equation
(13.5.1), it is more convenient to introduce the dissipation function or specific
power of dissipation, given by
(13.5.2)
where the dots designate time derivatives. The dissipation function is uniquely
determined by the strain rates (even for singular yield loci [10]), and we can
consequently write
(13.5.3)
bound theorem can now be stated as follows: The safety factor is the largest
statically admissible multiplier; i.e.,f?:. .\-.
A velocity field is called kinematically admissible if it satisfies the velocity
(or displacement) constraints and if the total external rate of work De done
by the actual loads on this velocity field is positive.
Let the generalized strain-rate vector associated with a given kinematically
admissible velocity field be designated by qf', where the asterisk is used to
indicate that this is not necessarily the true strain-rate vector but one that is
kinematically admissible. The internal dissipation function corresponding to
q'f can be determined from equation (13.5.3) and integrated over the complete
structure to obtain the total internal dissipation D 1• We now define a kine-
matically admissible multiplier .\ + as the ratio of the internal to the external
energy dissipations:
(13.5.4)
The upper bound theorem is then stated as follows: The safety factor is the
smallest kinematically admissible multiplier; i.e., f::;; .\ +. Both theorems can
be summarized by the relation
(13.5.5)
The above theorems furnish upper and lower bounds for the load. The
question of the uniqueness of the solution must still be considered. The
uniqueness of the safety factor follows from equation (13.5.5). It can also
be shown [4, 12, 13], using the principle of virtual work, that the stress field
at the start of plastic flow is unique except in the rigid regions and at stress
points which fall on the straight portions of yield loci (such as the Tresca
locus).
As a simple example of the use of these theorems, consider the indeter-
minate beam shown in Figure 13.5.1. The end A is built in and the end B is
A P 8
~! f f f f f !f f t f f !~~
(a)
(b)
simply supported and the distributed load is of intensity P. The beam will
collapse when yield hinges are formed at A and at some as yet unknown point
designated by x = g, The correct solution can readily be obtained. Let R be
the reaction at the point B. Then the moment at any x is given by
RL
r = Mo = 2 + 2v.r2 = 4.82843
PL2 -
p = Mo = 6 + 4V2 = 11.65685 (13.5.9)
- g . ;-
g= £ = 2 - v 2 = 0.58579
Equations (15.5.9) give the exact answer to the problem. However, for more
complicated problems the exact solution may become very difficult. An
approximate solution can be obtained by finding upper and lower bounds.
We note first that if some arbitrary value g were chosen for the hinge, an
upper bound would result, for the correct value of g will cause the yield
limit to be reached at a lower load. Let us then arbitrarily choose the mid-
point g = t as the hinge point. Then if the hinge A rotates through an
8, g will rotate through an angle 28 and the total internal work will be
The external load P moves an average distance of !L8 and the external work
Sec. 13-5] Theorems of Limit Analysis 317
PL2- 3M
4 - 0
(13.5.11)
PV = 12
or p+ =-
Mo
Comparing with equation (11.5.9), it is seen that an upper bound has been
found which is reasonably close.
A lower bound to P is obtained as follows. Calculating the moment
distribution due to the load P [assuming M(O) = - M 0 ] results in
(13.5.12)
For the above simple example there is little difference in the amount of
labor between the exact solution and the method of upper and lower bounds.
For more complex problems, however, the method of upper and lower bounds
is far superior.
318 Limit Analysis [Ch. 13
n = m- r (13.6.1)
l
4P
p
----------1
15 17 81
X~ I
I I
I I
I I
>I,
/
~I<~----2L------~~----2L
FIGURE 13.6.1 Two-bay frame.
Sec. 13-6] Method of Superposition of Mechanisms 319
2P
(a) (b)
(c)
FIGURE
1 (d)
basic or elementary mechanisms shown in Figure 13.6.2. There are three such
elementary mechanisms defined as follows:
I. Beam mechanisms as shown in (a) and (b). These are characterized by
one hinge at each end and one hinge under each concentrated load and one
hinge for a distributed load.
2. Frame or panel mechanisms as shown in (c). These are characterized
by hinges at the joints and supports but not under loads.
3. Joint mechanisms as shown in (d). These are fictitious mechanisms
representing the small rotation of a joint where three or more members are
connected.
320 Limit Analysis [Ch. 13
(13.6.2)
In the above equation the hinge at section 2 was assumed to take place in
the vertical leg, which is weaker, having a plastic moment of M 0 rather than
2M0 as in the beam.
For the mechanism of (b),
PL = (8£ - x)L M
. 2x(2L- x) 0
The frame can fail in one of the three mechanisms analyzed or in some
combination of them. The correct failure mechanism is the one requiring the
smallest load P. Hence in combining the mechanisms we choose only those
combinations which are likely to give smaller loads than the mechanisms
already available. Examination of the work equation (13.6.2) shows that to
reduce the load the work done by the load should be made as large as possible
and the internal work of the plastic moments should be made as small as
possible. This leads to the following rules for combining mechanisms:
Let us apply these rules to the present problem. The mechanism of Figure
13.6.2(c) is most likely to be correct, since it corresponds to the smallest value
of P. If it is combined with mode (a), the hinge at section 2 is eliminated, so
that by rule 1 the resultant value of P should be lower. This combined mode
is shown in Figure 13.6.3. The work equation can be obtained for this new
mechanism of Figure 13.6.3, but it is usually simpler to modify one of the
4P
~-----2L----~~~----2L----~~
previous work equations, here the one for the mechanism of Figure 13.6.2(c),
equation (13.6.5).
or PL = 2.75M0 (13.6.6)
This combined mechanism thus gives a lower collapse load and is therefore
more likely than any of the others. To determine whether it is correct, it is
necessary to check the other combinations. Neither mechanism (b) or (d)
of Figure 13.6.2 combines favorably with Figure 13.6.3, since no hinge is
thereby eliminated. However, both of these together will combine with
Figure 13.6.3 to eliminate the hinge of section 5 without introducing one at
10
//
~~~~----2L------~~-----2L------~
FIGURE 13.6.4 Combined mechanism.
+ 4P( ~ ) =
8
4P(L8) Mo(58) + 2M0 (68)
Hence
PL = 2.83M0 (13.6.7)
In the above equation the hinge at section 7 was assumed at the midpoint
of the beam according to rule 3. The correct value of x can be determined
from the condition that P should be a minimum. This gives a value of x equal
to 0.982L instead of L. The corresponding value of PL remains the same as
given by equation (13.6.7) to three significant figures.
Sec. 13-7] Limit Design 323
The mechanism yielding the lowest collapse load is the one of Figure
13.6.3, and this is therefore the correct collapse mode, the load being given
by equation (13.6.6). However, to verify that this is the correct solution, it is
necessary to calculate the remaining moments to make sure that none of
them exceeds the yield moment. The four independent equations of equilib-
rium are
2PL = 2M3 - M2 - M4
2PL = 2M7 - Ma - Ms
(13.6.8)
2PL = M 2 - M1 + M 5 - M 1o + M 9 - Ms
0 = M4 + Ms- Ma
Ma = 2Mo
Ms = -Mo
into equations (13.6.8), and solving for the remaining moments gives
Ma = -Mo (13.6.9)
Thus none of the remaining moments exceeds the fully plastic moment for
that member, so that the failure mode of figure 13.6.3 is the correct one.
The above example illustrates the use of the method of superposition of
mechanisms. Obviously the method requires some skill in combining favor-
able mechanisms. Such skill can readily be acquired by experience and
practice. The method then offers a relatively simple procedure for finding the
collapse load of framed structures under proportional loading.
where a1 corresponds to the ith element and the a1 are all specified, then the
design problem becomes one of finding the value of M 0 for collapse to occur
under the specified loads. This is called a restricted design problem [14]. To
find M 0 , the upper and lower bound theorems can be used as before, noting
that the lower bound theorem now gives an upper bound on the required value
of M 0 and the upper bound theorem gives a lower bound on the value of M 0 •
Thus in using the method of superposition of mechanisms of Section 13.6,
instead of combining mechanisms to obtain the lowest collapse load, we
combine the mechanisms to obtain the largest plastic moment. The procedures
for both types of problems are of course exactly the same.
For the illustrative problem of Section 13.6, as shown in Figure 13.6.1, the
highest value of the plastic moment M 0 is obtained for the combined mecha-
nism of Figure 13.6.3 and from equation (13.6.6) is given by
M0 = 0.364PL (13.7.1)
Problems
1. Consider a truss similar to the one shown in Figure 13.2.1, where the
between the vertical bar and each of the other two bars is 60°. Find
elastic and plastic safety factors if F = 20,000 lb.
References 325
References
1. G. Kazinczy, Experiments with Clamped Girders, Betonszemle, 2, Nos. 4,
5, and 6, 1914.
2. P. G. Hodge, Jr., Plastic Analysis of Structures, McGraw-Hill, New York,
1959.
3. A. A. Gvozdev. The Determination of the Value of the Collapse Load for
Statically Indeterminate Systems Undergoing Plastic Deformation, Proceed-
ings of the Conference on Plastic Deformations, Akademiia Nauk SSSR,
Moscow, 1938, pp. 19-33. Translated into English by R. l'v!. Haythorn-
thwaite, Intern. J. Mech. Sci., 1, 1960, pp. 322-355.
4. R. Hill, On the State of Stress in a Plastic-Rigid Body at the Yield Point,
Phil. Mag., 42, 1951, pp. 868-875.
5. R. Hill, A Note on Estimating the Yield-Point Loads in a Plastic-Rigid
Body, Phil. Mag., 43, 1952, pp. 353-355.
6. D. C. Drucker, H. J. Greenberg, and W. Prager, The Safety Factor for an
Elastic-Plastic Body in Plane Strain, J. Appl. Mech., 18, 1951, pp. 371-378.
7. D. C. Drucker, W. Prager, and H. J. Greenberg, Extended Limit Design
Theorems for Continuous Media, Quart. Appl. Math., 9, 1952, pp. 381-389.
8. H. J. Greenberg and W. Prager, Limit Design of Beams and Frames, Trans.
ASCE, 117, 1952, p. 447. First published as Tech. Rept. Al8-J, Brown Univ.
Press, Providence, R.I., 1949.
9. M. R. Horne, Fundamental Propositions in the Plastic Theory of Structures,
J. Inst. Civil Engrs. (London), 34, 1949-1950, pp. 174-177.
10. W. Prager, An Introduction to Plasticity, Addison-Wesley, Reading, Mass.,
1959.
11. D. C. Drucker, Limit Analysis and Design, Appl. Mech. Rev., 7, No. 10,
1954, pp. 421-423.
12. R. Hill, On the Problem of Uniqueness in the Theory of a Rigid-Plastic
Solid, J. Mech. Phys. Solids, 4, 1956, pp. 247-255; 5, 1957, pp, 1-8, 153-161,
302-307.
326 Limit Analysis [Ch. 13
General References
Baker, J. F., M. R. Horne, and J. Heyman, The Steel Skeleton, Vol. II, Cambridge
Univ. Press, Cambridge, 1956.
Commentary on Plastic Design in Steel, Progress Reports No. 1 and No. 2 of the
Joint WRC-ASCE Committee on Plasticity Related to Design, Proc. ASCE,
85, No. EM3, 1959.
Drucker, D. C., Plastic Design Methods, Advantages and Limitations, Brown
Univ; Div. of Appl. Math. Tech. Rept. No. 24, 1957.
Hodge, P. G., Jr., Plastic Analysis of Structures, McGraw-Hill, New York, 1959.
Massonnet, C. E., and M. A. Save, Plastic Analysis and Design, Vol. 1, Ginn
(Blaisdell), Boston, 1965.
Neal, B. G., The Plastic Methods of Structural Analysis, Wiley, New York,
2nd ed., 1963.
Van den Broek, J. A., Theory of Limit Design, Trans. ASCE, 105, 1940, pp.
638-661.
CHAPTER 14
CREEP
(14.1.1)
327
Creep [Ch. 14
328
Time
FIGURE 14.1.1 Typical uniaxial creep curve.
8 = 81 + 82
81 = tf:3t-2's (14.1.3)
82 = k
We see then that 81 is large for small times but becomes vanishingly small for
large times, whereas 82 is constant with time.
Many experiments on different materials have shown, however, that the
exponent i in equation (14.1.3) is not adequate and can vary between 0.4 and
0.85. Therefore, the relationship generally used is of the form
(14.1.4)
81 = (:3(a)tq (14.1.5)
e2 = k(a)t
Sec. 14-1] Basic Concepts 329
Equations (14.1. 5) imply that the curves for different stress levels are
geometrically similar, and this is approximately true.
In particular, the functions f3(a) and k(a) are frequently taken to be power
functions of the stress [2, 3]; i.e.,
Such expressions often fit the data fairly well. For example, for a gas turbine-
alloy steel (Allegheny 418, 12% Cr, 3% W) the data could be fitted fairly
well by assuming
- (61)1/q
t- II (14.1.9)
(14.1.10)
Thus the creep rate can be written as a function of stress, temperature, and
time, or of stress, temperature, and strain.
Basically, however, expressions of this type pose a fundamental difficulty.
For a relation of the form
8 =f(a, T, t, e) (14.1.11)
implies that there exists a "mechanical equation of state." That is, the creep
rate at any time depends on the state of the system at that time and is inde-
pendent of how or by what path the system got to this state. There is con-
clusive evidence that this can generally not be true but may be true under
330 Creep [Ch. 14
5 35,000 35,000
30,000
~--
25,000 25,000
I 20,000 20,000
0 20 40 60 80 100 120 140 0 20 40 60 80 100 120 140
Time, hr Time, hr
(a) (b)
0"2
Time
FIGURE 14.1.3 Transient effect on creep rate.
move to a point D which has a higher strain rate. The point Dis taken as the
point at which the strain is a1 /a 2 times the strain at A. The validity of this
type of approach requires much more experimental verification.
If one uses the time-hardening rule, then the creep rates at any stress are
given by equations (14.1.8). On the other hand, if the strain-hardening rule
is used, the creep rates are given by equations (14.1.10). We note that for
steady-state or secondary creep, all the rules become the same, since all the
stress curves are then parallel straight lines. (This is not evident from Figure
14.1.2, since the primary parts of the curves have been greatly exaggerated).
As mentioned before, all creep data are of the uniaxial type. How then do
we use this data for two-dimensional and three-dimensional problems? The
answer is, the same way as for any plasticity problem. Although the physical
processes involved in creep are undoubtedly different than in ordinary plastic
flow, we assume that the same relations hold for creep strains as for plastic
strains. For example, the Prandtl-Reuss relations can be used for computing
332 Creep [Ch. 14
the creep increments. Thus we assume an equivalent stress defined the same
way as in plasticity theory and an equivalent creep strain increment and write
etc. (14.2.1)
The computations are then performed as for any plasticity problem. Instead
of using the stress-strain curve, however, the creep curves, or equations such
as (14.1.8) or (14.1.10) representing these curves, are used. e0 and a. in these
equations are the equivalent strain rate and equivalent stress, respectively.
The successive-approximation method is very useful in making the calcula-
tions.
In general, solutions are obtained incrementally. One starts with a given
increment of time and solves the problem by successive approximations. The
next increment of time is then taken and the process repeated. The solution
can thus be extended to any time. For this purpose it is convenient to write
equations (14.1.8) as follows [using equation (14.1.6)]:
l::!ec = qBa':tq-l M
(14.2.2)
or l::!e 0 = Ka~ M
(14.2.3)
y=-c
FIGURE 14.3.1 Flat plate with temperature distribution.
(14.3.1)
(14.3.2)
(14.3.3)
from which
aX =E (
eX - aT - eX0 -
ile")
~
2 (14.3.4)
fc axdY =0
(14.3.5)
fc axydy = 0
334 Creep [Ch. 14
T = T 0 + 600(y 2 - t)
E = 28 x 10 6
(14.3. 7)
c= 1
a = 9.5 X 10-s
Substituting these values into equation (14.3.6) and noting the symmetry of
the problem results in
This equation, along with the stress-strain relation (14.3.4) and some relation
between stress and creep rate, are all that are needed to solve the plate
problem. Let it be assumed that the relation between stress and creep rate is
of the form
(14.3.9)
Note that in equation (14.3.9) the sign of ~e~ must be taken the same as the
sign of ax, as indicated. It will generally be necessary to write the creep law
in this form unless the stress exponent is an odd integer. The procedure for
obtaining the solution to this problem is now as follows:
1. At the start of the first time interval M, e~ is known to be zero and
~e~ is assumed to be zero. Substituting these values into (14.3.8) gives the
elastic solution as a first approximation to the total strains.
2. Substitute this first approximation for the strain distribution into the
stress-strain relation (14.3.4) and solve for the first approximation to the
stress distribution.
3. Substitute this first approximation to the stress distribution into the
creep relation (14.3.9) and solve for the second approximation to the incre-
mental creep strains during the first time interval.
Sec. 14-4] Creep in Rotating Disks 335
4. These incremental strains are substituted into equation (14.3.8) and the
iteration proceeds from equation (14.3.8) to (14.3.4) to (14.3.9) back to
(14.3.8), etc., until the procedure converges to the correct set of incremental
strains and stresses.
5. At the start of the second time increment, the total strains e~ are now
known and are equal to the incremental creep strains developed during the
previous time increment. In fact, the total creep strains at the beginning of
any time interval will always be known and will be equal to the accumulated
incremental strains up to that time interval. The procedure for calculating
the average stresses and strains for the second or any other time interval is
then the same as in steps 1, 2, 3, and 4.
The results of this calculation are presented in Figure 14.3.2. This figure
y =0 center
"'
Q.
shows a plot of the variation of the stress with time at the center and at the
edge of the plate. For this problem as well as the subsequent ones, the
unknown integrals were evaluated numerically using Simpson's rule.
The simple example of the fiat plate of Section 14.3 involves uniaxial stress,
making it possible to determine directly the creep strains once the total
stresses and strains are known. In most cases of practical interest the stresses
are biaxial or triaxial in nature. The general procedure is then the same as
discussed in Chapter 9. As an example we consider the case of creep in a
rotating disk.
336 Creep [Ch. 14
We assume that the von Mises yield criterion and the Prandtl-Reuss
stress-strain relations are valid for both plastic flow and creep, as indicated
by equation (14.2.1). For problems of thin disks with radial symmetry these
relations are
(14.4.1)
where as usual
(14.4.2)
(14.4.4)
where h is the thickness, p the density, and w the rotational speed. The com-
patibility equation in terms of stresses is derived in Section 9.5 and in the
notation used here is
-d (aa
dr E
- -E + a T + sao +
- JLar A
usac)
1 + p, a, - aa e~ - e~ !le~ - !leg
= ---y- --r-- + --r- + r
Sec. 14-4] Creep in Rotating Disks 337
Equations (14.4.4) and (14.4.5) must now be solved simultaneously for the
stresses ar and a 8 subject to the proper boundary conditions for the disk. At
any time t the total accumulated creep strains e~ and eg up to that time t are
known. The increments in creep strain lle~ and !leg for the next time interval
Llt are not yet known, but a set of values is assumed (such as the incremental
strains computed for the previous time interval). Since all the creep strains
now have known values, equations (14.4.4) and (14.4.5) form a linear pair of
equations which can readily be solved for the stresses ar and a 8 • From the
assumed values of lle~ and !leg, lle0 is calculated by the first of equations
(14.4.2). This value of flee corresponds to a particular value of a 8 as given by
equation (14.4.3). From these values of flee and a 8 as well as the stresses
computed from equations (14.4.4) and (14.4.5), new values for the creep
increments Lle~ and !leg are now computed from equations (14.4.1). These
better approximations to the incremental creep strains are put into equations
(14.4.4) and (14.4.5) and the process is repeated until convergence is obtained.
Equations (14.4.4) and (14.4.5) may be solved in a number of ways once
the incremental strains are assumed to be known. Two methods have been
described in Section 9.5, the finite-difference method and the integral-equation
method. The finite-difference method is generally preferable here, since the
disk thickness and other dimensions are changing with time. In connection
with this, it should be noted that since a disk under creep conditions will grow
and change dimensions with time, the values of r and h appearing in equations
(14.4.4) and (14.4.5) should be the true values at the time t and not the
original values at zero time. Thus if H is the original thickness at the
original radial position R, the current values h and r are given approximately
by [see equations (9.5.25) and (9.5.26)]
h= H
(1+ eg)(l + e~) (14.4.6)
r= R(l + eg)
by a function of stress times a function of time. The same problem will then
be treated including the elastic strains, to determine the effect of neglecting
the transient stress distribution upon the creep strains. The second case will
treat a similar disk using more complicated creep laws. The solution to the
problem will thus be presented using both the time-hardening and strain-
hardening rules.
(14.4.7)
(14.4.8)
(14.4.9)
Sec. 14-4] Creep in Rotating Disks 339
where Cis determined from the rim loading a,(b) and is given by
(14.4.10)
4. Calculate
l:le )l/6.2
ae = ( 4.41 X 10- 32 f:lt
2 3 4 5 6
Radius, in.
2 3 4 5 6
Radius, in.
figures that the agreement is excellent. The strains shown in Figure 14.4.2
were obtained by multiplying the strains computed using a M of 1 hour
by 180.
The same disk as in Case 1(a) was considered, with the elastic strains
included this time. The modulus of elasticity E was taken to be 18 x 10 6 ,
corresponding to a disk temperature of 1000°F. The creep rate was the same
as previously given for Case 1(a). To start with, a time interval M of0.01 hour
was chosen, and as a first approximation the creep increments l:leie and l:lefe
were all taken to be constant at 0.00001, the total creep strains e~ and eg of
course being zero at zero time. The calculation then proceeded as follows:
1. !:lee was computed from the relation
2. Equations (14.4.4) and (14.4.5) were solved for the stresses ar and a8
by the method described in Section 9.5.
3. a. was computed from the relation
Clie )1/6.2
ae = ( 4.41 X 10 32 l:lt
4. New approximations were obtained for the strain increments l:le~ and
!:leg from (14.4.1).
Sec. 14-4] Creep in Rotating Disks 341
5. Steps I through 4 were repeated until there was no change in two
successive computations of the strain increments.
The creep strains were thus computed at the end of O.OI hour. To obtain
the creep increments for the next O.OI hour the same procedure was followed,
except that the total creep strains e~ and eg were no longer zero but were
equal to the accumulated creep strains up to that time. A first approximation
to the creep increments <1e~ and <1eg was assumed (usually the values obtained
for the previous time interval), and steps I through 5 repeated. In this way
the incremental and total creep strains were computed up to I80 hours. The
time interval was arbitrarily chosen so that the stress during any interval
would not drop by more than approximately I,OOO psi. Thus this time
interval M was rapidly increased as the stress approached steady-state
conditions.
The results of this computation are shown in Figures I4.4.3 through I4.4.5.
Figure I4.4.3 shows the relaxation of the tangential stress with time at the
8. 40
vi
V>
"'
~ 30 _____________ __ _
:~_:::_:::_'='-=---~-='-:::::::
20 tangential stress
at rim
10
0.001
Time, hr
inner and outer radii of the disk. It is seen that steady-state conditions are
obtained at about 10 hours and that the tangential stress at the inner surface
persists for a long time at a slightly higher value than that computed neglecting
the transient stress distribution.
The tangential and axial creep strains at the end of 180 hours are shown
plotted in Figure I4.4.4 together with the scatter band of test results given in
reference [6]. From Figure I4.4.3 it can be seen that after a long period of
stability the stresses start rising. As the disk grows due to creep, the centrifugal
loading increases and the thickness at the center decreases, as can be seen
342 Creep [Ch. 14
- - including transient
conditions
~ scatter- bond test
results, Ref. (6)
2 3 4 5 6
Radius, in
i6.40[
e 6.20
<lJ
"0
".5
8 6.00 ----------------~
:~ 1.65~
2 1.45
<lJ
"0
·v;
E 1.25 ---------------~
.~ 1.00l
if
+-
0.90
£:~
f---Vl
·"o.so L---~-L-LJ-LILllui~~----L-~1~1-Lt~l~t~l~t~I~~--J-J-~I~I~I~tl
1 10 100 1,000
Time, hr
from Figure 14.4.5. This increase in centrifugal loading and decrease in cross
section will eventually balance and finally surpass the stress relaxation, due
to the creep. The stresses will start increasing all over the disk and the disk
may start creeping at an accelerating rate. This is illustrated in Figure 14.4.3,
Sec. 14-4] Creep in Rotating Disks 343
which shows that after a long period of stability the stresses gradually start
rising. The disk has now reached an unstable condition which may result in
failure.
In the previous examples the creep curves were assumed linear with the
creep rate independent of time. It will now be shown that more complicated
creep curves and creep laws can be used without any appreciable increase in
the complexity of the computations. Any nonlinear creep law can be used;
for illustration it will be assumed that the creep data can be represented by
(14.4.11)
The same disk as in case 1 will now be considered, with the creep data given
by equation (4.4.11 ), using both the time-hardening and the strain-hardening
rules previously mentioned.
For the time-hardening rule, since the creep rate depends on the actual
time elapsed, the creep rate at constant stress at the middle of the time
interval between t and t + l1t is obtained directly from (14.4.11):
!J.t)q -1
Ae0 = qBa~ (t +2 At (14.4.12)
where ae is the stress at the middle of the time interval and is assumed to be
constant during the interval. Solving (14.4.12) for a. gives
= (~)1/m ( !J.t)<l-q)/m
ae qB !J.t t + 2
(14.4.13)
For the strain-hardening rule, the creep rate depends upon the total
accumulated creep rather than the accumulated time. Eliminating the time
between equations ( 14.4.11) and (14.4.12) results in
where e0 is the total accumulated creep strain up to the beginning of the time
interval under consideration and Lle0 is the strain increment during the time
interval. The solution to this problem is now obtained in exactly the same
way as for case l(b), except that in step 3, a. is computed by equation (14.4.13)
344 Creep [Ch. 14
for the time-hardening rule and by equation (14.4.14) for the strain-hardening
rule. For the time-hardening rule, track must be kept of the total elapsed
time, whereas for the strain-hardening rule, track must be kept of the total
accumulated strain.
Part of the results of this calculation for m equal to 6, q equal to ~' and
B equal to 1.5 x 10-ao are presented in Figures 14.4.6 and 14.4.7.
The creep curves represented by these constants are shown in Figure
14.4.8. As would be expected, Figure 14.4.6 shows that the stress does not
relax quite as rapidly using the strain-hardening rule as with the time-
hardening rule, since the creep rates will generally be lower for the strain-
hardening rule. This can be used to explain the results of Figure 14.4.7, which
- - time hardening
----strain hardening
40 .... ,.....
------ ---
'-.-. tangential stress at bore
·u;
-----------
0.
Time, hr
Time, hr
120 140
Time, hr
shows that the tangential creep strain is essentially the same for both rules.
(Although the strain-hardening rule actually gave slightly lower creep strains,
the difference is too small to show up on the figure.) Although as seen from
Figure 14.1.2, in going to a given stress the creep rate will be smaller for the
strain-hardening rule than for the time-hardening rule, this is compensated
for in this problem by the fact that the stress does not relax as fast, so that
the total creep turns out to be about the same using either rule.
A comprehensive discussion of creep in rotating disks, including the effects
of the transient period, can be found in a series of papers by Wahl [6, 7, 8, 9,
10].
References
1. E. N. da C. Andrade, On the Viscous.Flow of Metal and Allied Phenomena,
Proc. Roy. Soc. (London), A84, 1910, p. 1.
2. F. H. Norton, Creep of Steel at High Temperatures, McGraw-Hill, New
York, 1929, p. 67.
3. R. W. Bailey, The Utilization of Creep Test Data in Engineering Design,
Proc. Jnst. Mech. Engrs., 131, London, 1935, pp. 186-205, 260-265.
4. Yu. N. Rabotnov, On the Equation of State of Creep, Proceedings of the
International Conference on Creep, Vol. 2, New York-London, 1963, p. 117.
346 Creep [Ch. 14
General References
Arutyunyan, N. Kh., Some Problems in the Theory of Creep, translated by H. E.
Nowottny, Pergamon Press, London, 1966.
Finnie, J., and W. R. Heller, Creep of Engineering Materials, McGraw-Hill,
New York, 1959.
Hult, J., Creep in Engineering Structures, Ginn (Blaisdell), Boston, 1966.
Odquist, F. K. G., Mathematical Theory of Creep and Creep Rupture, Oxford
Univ. Press, London, 1966.
INDEX
347
348 Index
Johnson, W., 22, 23, 134, 163, 258, Marin, J., 11, 22, 97
259 Martin, H. C., 233
Massonnet, C. E., 326
Kantorovich, V., 258 Maximum load point, 6, 8, 9, 11
Kazinczy, G., 301, 325 Maximum shear theory, 73
Kinematically admissible multiplier, Maximum strain energy theory, 74
315 Maximum strain theory, 72
Kinematically admissible velocity field, Maximum stress theory, 71
284, 315 Mechanism(s), 30
Kinematic hardening, 95 elementary, 318
Kinematic mechanism, 310 beam, 319
Kobayashi, S., 299 frame, 319
Koff, W., 97 joint, 319
Koiter, W. T., 134, 156, 163 panel, 319
Korn, G. A., 299 portal, 320
Korn, T. M., 299 fictitious, 319
Kronecker delta, 27 linearly independent, 318
Krylov, V. 1., 258 rules for combining, 321
Melan, E., 119, 133
Mellor, P. B., 22, 23, 134, 163, 258,
Lee, E. H., 163 259
Levy, M., 1, 3, 100, 133 Membrane analogy, 245
Levy-Mises equations, 100, 261 Membrane-roof analogy, 247
Life-fraction law, 330-31 Mendelson, A., 134, 212, 233, 346
Limit analysis, 300-26 Metal forming processes, 2, 12, 15, 260
of beams, 305-310 Mikhlin, S. G., 212
with concentrated loads, 307-310 Minimum weight design, 324
in pure bending, 305-307 Mises, R. von, 1, 3, 77, 100, 133
design of structures, 300-301 Mises, von, criterion, 75, 115, 129
of frames, 310-12 flow rule, 100-1 04
of simple truss, 301-305 stress-strain relation, 100-104
theorems, 312-17 Mises, von, ellipse, 76
lower bound, 313-15 Mohr, 0., 43
upper bound, 313-15 Mohr's circle(s), 37-39, 280
Limit design, 301, 323, 324 for plastic strain increments, 101
restricted design problem, 324 pole of, 280
Livesley, R. K., 324, 326 for stress, 101
Loading, definition of, 93 Mohr's diagram, see Mohr's circles
Loading function, 92, 112 Mohr's theory, 77
Load point, maximum, 6, 8, 9, 11 Multiplier
Lode, W., 88, 89, 97, 109 kinematically admissible, 315
Lode's statically admissible 314, 315
strain parameter, 109 Muskhelishvili, N. I., 226, 233
stress parameter, 88
Lower bound theorem, 285, 313-15 Nadai, A., 258
Lower yield point, 6 Naghdi, P.M., 97, 134
Lubahn, J.D., 22 Navier equations, 68
Ludwik, P., 7, 20, 22 Neal, B. G., 318, 326
Necking, 6, 9, 12
Mclean, D., 22 Neuber, H., 224, 233
Magnusson, A. W., 22 Neutral loading, 93
Manjoine, M. J., 15, 22, 346 Norton, F. H., 345
Manson,S.S. 134,212,233,346 Novozhilov, V. V., 63
350 Index