COLLEGE OF PETROLEUM & MINERALS
LIBRARY
Marvin J. Forray
Professor of Mathematics
C. W. Post College of Long Island University
Variational
Caleulus
in Science
and Engineering
McGRAW-HILL BOOK COMPANY )
New York San Francisco Toronto London Sydneyyy. MINERALS,
GE OF PETROLEY!
ms LIBRARY
my
we:
On
VARIATIONAL CALCULUS IN SCIENCE AND ENGINEERING
Copyright © 1968 by McGraw-Hill, Inc. All Rights Reserved.
Printed in the United States of America. No part of this publication
may be reproduced, stored in a retrieval system, or transmitted,
in any form or by any means, electronic, mechanical, photocopy-
ing, recording, or otherwise, without the prior written permission
of the publisher. Library of Congress Catalog Card Number 68-20718
21584
1234567890 MAMM 7543210698Preface
The calculus of variations is assuming an increasingly important role
in the fields of analysis, physics, and engineering. While the subject
originated as a study of certain isolated maximum and minimum problems
not treatable by the techniques of elementary calculus, it is at present
a powerful method for the solution of problems in dynamics and statics
of rigid bodies, general elasticity (including thermal effects), the theory
of plates and shells, vibrations, optics, quantum mechanics, optimiza-
tion of orbits and controls, ete.
Students of mathematics, engineering, and physics receive in many
instances only fragmentary instruction in the method and application of
the variational calculus. In his quest for additional understanding of
this subject and to enhance his ability to read current articles the student
usually turns to mathematics books on the variational calculus. While
there are some excellent texts in this area, they are difficult to absorb,
particularly by the average nonmathematician. This book is written
in the hope that it will make the subject more palatable without sacrifice
of rigor. Methods will be developed and extensively illustrated with
worked examples. Over twenty years of teaching experience on the high
school, college, and graduate levels has taught the author that the best,
way to make a subject “stick” is to illustrate the method profusely,
because—unfortunately—abstract concepts are quickly forgotten or
misapplied.
33980vi Preface
The first three sections of Chapter 1 are relatively elementary and
should serve as a review of the elements of maximum and minimum for
functions of one and more than one variable. A number of illustrations
should help the reader to renew his acquaintance with these basic calculus
concepts, which are significant in the variational calculus also. Sections
1.4 and 1.5 serve as an introduction to extremum problems with con-
straints using the multiplier method of Lagrange. This is done at a
leisurely pace because it is anticipated that part or perhaps all of this
may be new to some readers.
Chapter 2 begins by introducing the concept of the functional and in
particular what is meant by the first variation of a functional. The
variational calculus is then treated formalistically to show the reader the
potency of this procedure. In particular, the problems of the cantilever
beam and circular plate are exhibited. Many of the concepts which
are going to be justified in detail, such as the fundamental lemma of
the variational calculus and the commutativity of the operators d/dz
and 6, are utilized. The main objective is for us to see the forest before
we start clearing the trees. This chapter closes with a few historical
remarks and a derivation of some significant classical examples including
the isoperimetric, brachistochrone, shortest distance, and minimum sur-
face of revolution problems.
Chapter 3 is initially devoted to the basic problem of formulating the
Euler differential equation for a wide variety of problems. In con-
junction with this, the basic lemma of the variational calculus is estab-
lished for both one- and two-dimensional problems. Both mathematical
and physical examples are given to illustrate the formulation of the Euler
equations. The solutions of Euler’s differential equations for various
cases are treated in detail. Natiiral boundary conditions followed by
a systematic treatment of the variation is given next. Applications are
presented in both of these areas. The remainder of the chapter treats
the Lagrange multiplier method for functionals.
In Chapter 4 we have the application of the variational calculus to
discrete mechanics. In particular, starting with Newton’s law of motion,
we deduce Hamilton’s principle for a single particle; and for a conserva-
tive system, the Lagrangian function is introduced. This is followed by
a generalization to n variables, and Lagrange’s equations are deduced. A
number of examples and a detailed analysis of free vibrations of a con-
servative system about a position of equilibrium are given. This is
illustrated in particular by the vibration of a light string with two con-
centrated masses attached.
Chapter 5 treats the elements of the theory of elasticity. The concepts
of displacement, strain, and stress are presented in some detail including
equations of equilibrium and compatibility, Also, the transformationsof components of strain and stress tensors about a point are deduced and
formulas for principal strains and stresses are obtained. Both Hooke’s
law and its generalization are developed and, furthermore, the significance
of the elastic constants for isotropic materials is demonstrated. Then
the important concept of strain energy is systematically deduced and a
proof is given that, in general, the strain energy density is a positive
definite form of the strain components.
Chapter 6 is, in a sense, a continuation of Chapter 5, because it starts
with the proof of the fact that the potential energy of a static elastic
continuum must be stationary. This result is then applied to three
significant problems, namely, the static response of an elastic string,
a beam column, and the deduction of the three-dimensional displacement
equations of elasticity. From the concept of stationary potential energy
we deduce Hamilton’s principle for a dynamic elastic continuum and
illustrate it with the flexural vibrations of elastic strings and beams. At
this point, it is expedient to introduce indicial notation and to apply it to a
general form of Castigliano’s least-work principle. This theorem, or its
corollaries, yields the true stress field from the many states of stress which
satisfy the differential equations of equilibrium and traction boundary
conditions. They result in the additional requirement that the com-
patibility equations must be satisfied by the actual stress field and,
furthermore, the complementary energy must be stationary. In the next
section a variational theorem due to E. Reissner simultaneously yields
relationships between the stress and displacement components, the
equilibrium equations, and the boundary conditions for small (or infini-
tesimal) displacements. This is followed by a generalization of a second
theorem by Castigliano which directly gives the displacements and
rotations at arbitrary points of an elastic body. The Maxwell law of
reciprocity is deduced for linear elastic bodies and exemplified for a
cantilever beam.
Chapters 7 and 8 treat some of the direct methods of the calculus of
variations. In Chapter 7, the Rayleigh-Ritz method is first outlined and
then exemplified by the solution of a particular differential equation and
the bending of a clamped rectangular plate. Then the method is applied
to simple vibration and stability problems. The torsion of circular
cyl ndrical shafts and more general prismatic bodies follow in some detail.
This problem is then recast in variational form and exemplified by the
torsion of elliptic and rectangular cylinders.
In the last chapter, the methods of Galerkin, Kantorovich, and Euler
are developed. After the Galerkin method is introduced, its equivalence
with the Rayleigh-Ritz method is established for self-adjoint differential
equations of order two. Further examples of this equivalence are given
for the functionals associated with the bending of plates and membranes.viii Preface
This is followed by theCingeniows method of Kantorovich which serves
as a generalization of the Rayleigh-Ritz procedure. Also, the method
is illustrated for the bending of a clamped rectangular plate. The next
section is devoted to the Buler method of finite differences. A detailed
application is given to the determination of an extremum associated with
a particular functional. In the last section, examples of problems
without solutions are demonstrated. Many illustrations of “plausible”
extrema problems for which there are no solutions may be given, thereby
revealing the significance of questions of existence (and uniqueness).
The Appendix is devoted to a more extensive treatment of the summa-
tion convention and Cartesian tensors than is presented in the body of the
book. In particular, the reader should find this appendix most helpful
in his study of Sections 6.3 to 6.6 of Chapter 6.
The author is indebted to Professor Robert V. Iosue for critically
reading the manuscript and making many useful suggestions for improve-
ment of the text. Also, it is a pleasure to acknowledge the very skillful
typing assistance of Mrs. Rosalie Ciaravino, Mrs. Doris Vandereedt, and
Mrs. Stasia Polster in different phases of the manuscript development.
To his wife Muriel, the author is indebted for her deep affection, direct
help and persevering efforts in encouraging the completion of this task.
Therefore, this book is affectionately dedicated to her.
Marvin J. ForayContents
Preface ov
ONE Maximaand Minima 1
1 Introduction 1
2 Necessary and Sufficient Conditions 2
3 Illustrations 6
4 Mazimum and Minimum Problems with Constraints 10
5 Lagrange Multiplier Method—General Case 14
TWO = Introductory Problems of the Variational
Caleulus 18
1 Concept of Functionals 18
2 Stationary Values for Functionals 20
3 Classica! Problems of the Variational Calculus 27
THREE Euler-Lagrange Development with
Applications 31
1 The Basic Problem—Euler's Equation 32
2 Euler Equation for More General Cases 35
3. Solution of Euler's Differential Equations 44enrane
Contents
Natural Boundary Conditions 49
The Variational Notation 58
Lagrange Multipliers for Funetionals 62
Isoperimetric Problems 65
Isoperimetric Problems—Generalization 70
FOUR Hamilton’s Principle and Lagrange’s
ansere
Equations 75
Hamilton's Principle for a Single Particle 76
Degrees of Freedom—Generalized Coordinates 80
Hamilton's Principle for a System of n Parlicles 81
Lagrange's Equations with Applications 83
Lagrange's Equations for a General System 87
Normal Oscillations 90
FIVE Deformable Bodies—Theory of Elasticity 95
BE CerAaneeNnH
SIX
NAnR ENE
aunewe
VEN Rayleigh-i
Strain 96
Transformation of Strain Components—Principal Azes 100
Theory of Small Displacements 104
Stress 107
Relation among Traction Vectors at a Point 110
Equilibrium Equations 112
Principal Stress at a Point 114
Hooke's Law 115
Generalized Hooke's Law 117
Significance of Elastic Constants for Isotropic Materials 119
Strain Energy 123
Energy Principles, Methods, and
Applications 126
Theorem of Minimum Potential Energy 126
Hamilton's Principle for an Elastic Continuum 134
Indicial Notation—Complementary Energy Density 137
General Form of Castigliano’s Least-work Theorem 144
Reissner's Variational Theorem of Three-dimensional Elasticity 147
A General Form of Castigliano’s Theorem on Deflections 149
Principle of Superposition and Mazwell’s Reciprocity Law 154
itz Method 157
An Outline of the Rayleigh-Ritz Method 158
Rayleigh-Ritz Method Applied to Vibration and Stability Problems 164
Torsion of Circular Cylindrical Shafts 174
Torsion of Prismatic Bodies 174
Transformation to a Dirichlet Problem—The Stress Function 177
Variational Form of Saint-Venant’s Problem, 182EIGHT Methods of Galerkin, Kantorovich, and
Euler 189
Galerkin Method 189
Equivalence of Rayleigh-Ritz and Galerkin Methods 193
Method of Kantorovich 199
Euler Method of Finite Differences 204
Final Remarks—Problems without Solutions 204
asewe
APPENDIX Summation Convention and Cartesian
Tensors 207
1 Summation Convention 207
2 Cartesian Tensors 240
Bibliography 213
Author Index 245
Subject Index 217ONE
Maxima and Minima
1 Introduction
The calculus of variations is concerned primarily with the determina-
tion of maxima and minima of certain expressions involving unknown
functions. Certain techniques and concepts of the differential calculus
are applicable to the variational calculus. ‘These will be reviewed briefly.
One of the basic theorems of the calculus (due to Weierstrass) is that
if a function f(z) is defined and continuous at each point of a finite closed
interval a <2
0] at
that point. If f’’(xo) = 0 and successive differentiation shows that the
first derivative which is not zero at z = 2» is odd, no maximum or mini-
mum occurs. If, on the other hand, it is even-ordered, a maximum occurs
when the value is negative, and a minimum when positive.
The problem of determining stationary values for a function of more
than one variable is more involved. The theorem of Weierstrass still
applies, namely, every function which is continuous in a closed domain D
of the variables possesses a largest and a smallest value in the interior or
on the boundary of the region. Suppose u = f(x1,22, . . . ,2a) is differ-
entiable in a domain D; then a necessary condition for an extremum at a
point P in the interior is
id ee
dagen “379 (2.1)
at P; that is, df =0.! The sufficiency conditions are more compli-
cated. In the case of a function of two variables, the function f(x,y)
yields a maximum at a point P if, in addition to f, =f, = 0 atP,
See <0 and fecfy — fay? > 0. It results in a minimum when f.:>0
and fzzfy —fxa? > 0. If, on the other hand, fesfyy — fey? <0, the
stationary value is neither a maximum nor a minimum. The case
Sexy — fy? = 0 remains undecided; that is, if it occurs, an alternate
procedure must be employed.
The above conditions have a simple geometrical interpretation. In
the case of a function of two variables, the necessary conditions for a
stationary value, fz = f, = 0 at (co,yo), imply that the tangent plane to
the surface z = f(x,y) at this point is horizontal (i.e., parallel to the
zy plane). If the point is an extremum (maximum or minimum), then,
1 The quantity df = > 42, dz; by definition. Note that, for convenience, we will
a
often use subscripts to denote partial differentiation with respect to that variable.
Thus fz = af/ar and fey = a*f/az dy, ete.in its neighborhood, the tangent plane does not intersect the surface.
In the case of a saddle point (no maximum or minimum) the plane cuts
the surface in a curve which has several branches at the point.
‘The sufficient condition for a maximum or minimum in the case of
functions of more than one variable is expressed most conveniently in
terms of quadratic form-matrix notation, which we will treat next.
DEFINITION: A quadratic form is a homogeneous polynomial of
the second degree in several variables. A quadratic form! may be
represented in the following form
F(eit2, . . 2m) = Guta + Greet, + + + Gintstn
+ autor, + Arete? + + A Aant2tn
t+ Gnitnti + Onatnte + * + Gnntn®
where without loss of generality, ai; = aj The matrix
ay a2 in
an a aay
Ae ee (2.2)
On Mascaras Onn
is called the matrix of the quadratic form. Because of the manner in
which it was composed, A is a symmetric matrix; that is, a, = a ‘Thus,
every quadratic form is associated in a natural way with a unique sym-
metric matrix and, conversely, every symmetric matrix may be associated
with a certain quadratic form.
A quadratic form may be written in an abbreviated matrix notation as
follows:
F(@ija, . «+ sn) = B(Gut1 + Grate + + + + + Ginn)
+ we(derti + Geeta + + + + + Genta)
ee es
+ &n(Qniti + dnote + + + Gnntn)
= [tr 2 ++ ta) (Ques + ait, + + dante
Gaye, + Arte + + + + + Aonte
Gnr%1 + On2t2 + °° * + Anat
=f me +++ tel fon ai ais --- Gu] [a
@o1 G22 Ges °° * Gon x
ni Gn2 Gna ~ °° nn. Tn,
F(tyta, .. . te) = 2! Ax (2.3)
1 The letters au, aiz, . . . » nn are taken to represent arbitrary real numbers.4 One
where 2’ = [21 22 * ~*~ a] is a row matrix which is the transpose of
the column matrix
Zn
A real quadratic form is called positive definite if, for real values of the
variables, its value is always positive, with the obvious exception of
y= t= =a, =0.
Example 1: 22+ m?+ 2+ -- + +? is a positive definite
form.
The term “positive definite” is also extended to symmetric matrices.
Areal symmetric matriz [A] = [ay] is called positive definite if the quadratic
form
2
Y aerers
1 it
F(a1,t2,03 - - » stn) =
ime
is positive definite. Thus, for example, the unit matrix is positive
definite, since the quadratic form corresponding to it is positive definite.
The following theorem (due to Sylvester) is significant for determining
necessary and sufficient conditions for maxima and minima.
THEOREM: A quadratic form )' )) airs; is positive definite if and
Aah
only if all the determinants (principal first minors)
ay a2
Ar = |an| 42 = aos
An =
are positive.
Similarly, a quadratic form F is negative definite if —F is positive
definite, while a matrix [A] is negative definite if [— A] is positive definite.
If a form is neither positive or negative definite, then it is said to be
indefinite.
Example 2: The quadratic form
P= ay? + 2x, + 5x3? — Qayte + 4aits — deersis easily shown to be positive definite. For this example,
eae eee 2
faJ =|-1 2 -2
2 -2 5
so that
4: = [1] =
1
aa-|_}
1
A= |-1
2
Since all the principal first minors are positive, it follows that F is positive
definite. In fact, the quadratic form can be written equivalently as
= (ey — a2 + 2s)? + a? + as?
which, being the sum of squares, is zero if and only if
tm —t%+27=0 w%=0 and 2s=0
which implies that 21 = 22 = x3 = 0.
On the other hand, the quadratic form
P = 2x,’ + 2ae? + 2xs? — 4a, — Avors + ear
1 -2 1| (x1
[z. we wal] -2 2 —2| 4m
1 -2 2) lea,
is readily shown to be indefinite. The principal minors are given by
A= [1] =1
1 -2
oe
et
4s=|-2 2 -2
1-2 2
Two particular values of F are F(1,0,0) = 1 and F(1,1,1) = —1, thereby
verifying that the form is actually indefinite.
Let f(e1,2n3, . . . ;2n) be three times continuously differentiable in
the neighborhood of a stationary point P: 2 = 2° @ = 1, 2,.... ,
That is, f2, = 7°, = =f’, =0. Consider the second total differ-6 One
ential of f at the point 2°; then df? = )' ) f0.,, dx dz; This is a
afr
quadratic form in the variables dz, . . . ,dtn. The quantity d’f? may be
(1) Positive definite
(2) Negative definite
(3) Indefinite
It can be shown by Taylor’s theorem that, corresponding to (1), f has a
minimum, and to (2), a maximum, whereas condition (3) yields neither a
maximum nor a minimum provided that
Fin. Seer °° * Son
D=lfie fen *** Sem| 9
3. Illustrations
Example 1: Find the absolute maximum and minimum for
f@@)=x'-2 -2<2<2
Solution: f'(x) = 4° — 5x4
For critical values, f’(x) = 0 must necessarily hold. Therefore,
2(4 — 52) =0
z = 0 (3 times)
and
cat
are the critical values.
S'(@) = 12a? — 202%
F'(8) = 12(8)? — 20(8)* = ($)2(12 — 16) < 0
Thus x = yields a local or relative maximum, f($) = 44/5%. Also,
J) = 0, so that higher derivatives must be calculated and evaluated
atz = 0,
"(@) = Az — 602% 70) = 0
f"(@) = 24-1202 f”"(0) = 24
Since the first derivative which does not vanish is even-ordered and its
value is positive, we have a relative minimum at z = 0, f(0) = 0.
Aref($) = 44/5* and f(0) = 0 the absolute maximum and minimum for
f(x) in —2 0.
Comparison of (@j) and (i) results in u |" = «| 6
Example 3: Find the dimensions of an open rectangular box with a
capacity of 4 cu ft which shall have the smallest possible surface area.
Solution: Let x and y be the horizontal edges and z the depth, all
positive from physical considerations. The surface area is
A = ay + nz + 2ye
The volume is ayz = 4, so that z = 4/zy. Therefore,
. ee
Aamtits
For critical values
8
~ a
Ar=8 One
so that
r=y=2 and A=12
‘The corresponding depth, z = 1, is half the edge.
‘The second derivatives at the point (2,2) are
16 16
An => Ay = qt? daa
so that at (2,2)
AgAy — Ax’ =
Since Az: > 0, we have arelative minimum. This may be shown to be an
absolute minimum.
Example 4: Test for relative maxima and minima:
flaye) = 22 + y® + 382? — ay + 2a2 + ye
Solution:
So = 2 —y +2
fy=%yr-aote
fe = 62+ 2 +y
For critical values, f2 = fy = fe = 0, which yields 2 = y = 2 = 0.
For the determination of whether we have a maximum or minimum, the
principal minor test will be used.
fos =2>0°
fox fv else eee _
tee ‘| “| | ~870
Jez fav Sue 22
Juz Sv a7 oe ed 0
fee fav oie
Thus
I(x,yz) = F,0,0) = 0
in the neighborhood of the point; that is, (0,0,0) yields a relative
minimum."
Example 5: Snell’s Law of Refraction. A very significant applica-
tion of maxima and minima is to the derivation of Snell’s law of refraction.
It is based upon a minimum principle of Fermat which states that the time
0, y > 0).
The area of the rectangle is f(x,y) = zy, and f is to be maximized subject
to the restriction that the perimeter is L. Thus the constraintis
g(a,y) = 2x + 2y-L=0
at (a,b) (4.1)
(4.2)12 One
Substitution of the expressions for f(z,y) and g(z,y) into (4.2) yields
y+2r=0 @
z+2=0 Gi)
Also,
Qe +24 —L=0 Gi)
‘The simultaneous solution of (i) to (lii) results in
zea yad=% md y=-2
Therefore the square (a = b) makes the area stationary for a given perim-
eter. Furthermore, it is easily shown that the square actually maximizes
the area, It is also of interest that this example has a “dual” result;
namely, that of all rectangles of given area the square has the least
perimeter.
‘The above argument may fail to be valid if the curve g(x,y) = 0 has
a singular point (say, a cusp) at ¢ = a, y = 6. By a singular point we
mean a point for which both partial derivatives are zero. For example,
the origin (0,0) is a singular point of the curve y? — 2* = 0 which actually
has a cusp there.
In any event, our intuitive investigation suggests the following rule
(which will be established rigorously next):
In order that a function f(2,y) attain an extremum at =a, y =b
subject to the restriction g(x,y) = 0 where gs, = gy],, = 0 does not
occur, it is necessary that a constant of proportionality \ exists such that
Fe(a,b) + dgo(a,b) = 0
Fu(a,b) + rg,(a,b) = 0
are satisfied together with the equation g(x,y) = 0. Theabove rule is
known as Lagrange’s method of undetermined multipliers, and the factor
is known as Lagrange’s multiplier.
As expected, the analytic proof of the validity of the multiplier rule
is established by reducing the constraint problem to the known problem
with “free” extreme values. Let it be assumed that at the extreme point
(ab) at least one of the partial derivatives, say g,(a,b), does not equal zero.
Then, from the implicit function theorem of the calculus {assuming that
9(z,y) has continuous first partial derivatives] it is known that there exists
a neighborhood of (a,b) for which the equation g =0 determines y
uniquely as a continuously differentiable function of z, say y = ¥(z).
Substitution of y = ¥(z) into f(2,y) results in the function f(x,¥(z)), which
must have a free, i.e., unconstrained, stationary value at the point x = a.Therefore,
'@) =f2+ iW @) =0 (4.3a)
must hold atx =a, Also, from g(x,y) = g(z,¥(2)) = 0 we have
go: + gW'(z) = 0 (4.36)
Multiplication of (4.3b) by \ = —f,/g, and addition of this to (4.3a)
yields
fet dge + (fy + Agu)/(z) = 0 (4.4)
But by our choice of 2,
fy t dy =0 (4.5)
and this reduces (4.4) to
foto: =0 (4.6)
Equations (4.5) and (4.6) are precisely those of the method of undeter-
mined multipliers, and therefore the procedure is established.
The multiplier method can be summarized by the rule:
V Form F(z,y,d) = f(v,y) + dg(@,y) and treat the variables z, y, and ) as
if they are independent of each other. Then make F stationary by setting
the partial derivatives with respect to each of the variables equal to
zero, i.e,
oF
ag Oth tM
fit
This is identical with (4.5), (4.6), and the constraint condition.
Finally, before generalizing our results, we shall indicate the application
of differentials to this problem. In order that the function f(z,y) may
have an extreme value at the point z = a, y = b subject to the subsidiary
equation g(x,y) = 0, it is necessary that the differential df shall vanish at
that point; ie,
ee a7
Also from g(z,y) = 0,
Y/Y 9x(ajb) dz + 9,(a,b) dy = 0 48)
Multiplication of (4.8) by \ and addition to (4.7) yields
[f-(a,b) + dge(a,b)] dz + [f,(a,b) + dg,(a,0)] dy = 014 One
Choose ) so that
Sy(a,b) + g,(ab) = 0 assuming g,(a,b) + 0
Then since dz # 0 in general,
Sela,b) + dgz(a,b) = 0
and Eqs. (4.5) and (4.6) have been deduced again.
5 Lagrange Multiplier Method—General Case
More generally, consider the problem of extremizing a function of n
variables
F(wiyte, . . - sn) (5.1)
subject to constraints
gr(ti,22, .. . tn) = 0
9r(ti,G2, . . . stn)
(5.2)
of the n quai terms of the remaining n — k variables. Substitu-
tion into (5.1) yields a function of n — k independent variables. This,
of course, may then be treated as an ordinary maximum and minimum
problem by using the techniques previously developed.
Alternatively, the following symmetric procedure may be employed.
Since f is to be made stationary, it follows that
g= VS Da, = (5.3)
a
However, the n variables are not independent, but instead are subject to
the constraints (5.2), so that only n — k variables are independent. The
differentials dz; must satisfy k conditions of constraint, that is,
995 ae = ee
3 Ben 0 j=1,2,...,k (5.4)
Multiply (5.4) by A; and sum with respect to j from j = 1 to j =k.
Combination with (5.3) yields
d ¢ + i #) dx; = 0 (5.5)
which is true for arbitrary \1, Xs, . . . , Xe and arbitrary values of n — kof the differentials. In expanded form we have
af i Ogata
z the tg to +g ms)
+ “(zt +S Se ts #) des
foe + (Zang aye “+n 32) dn =0 (5.6)
Choose the )’s so that k of the n bracketed expressions are zero. It then
follows that the remaining n — k bracketed expressions must be zero,
since the n — k remaining differentials are arbitrary. Thus for a relative
maximum or minimum, we have
oa e
— gia 0 G=1,2...50 Co)
+e. t=O §1,2,...,k
as necessary conditions. This system (5.7) is n + k equations to be
solved for the same number of unknowns 21, 2; . . « ;tnjMyAz . » 0 y Me
These equations (5.7) are also obtained by forming
F(@it2, . «+ s®nyAayd2, - «+ ade)
k
= ferry, ... tn) + Y Agilent, - - stn)
j=
and setting the partial derivatives with respect to the n +k variables
2%... Tn My bay». . » Xe Equal to zero:
oe
Ox;
(58)
Fo 5=1,2...,k
ay 12, ,
This again is the Lagrangian rule.
Example 1: Minimize x? + 2:? + - - + + 24? subject to the condi-
tion ayer + dete + °° + ante = 1, where clearly a:* + a2? +--+
a,? is positive.
Solution: Form
F(a1,22, « « - jn)
Satta tos tat + Mati + ante + + ~ + ante — 1)
oF
Ge, 7 2 tm = 0
SF = 20, + dar = 0
Ooze
oF x 2m, + da = 0
at,16 One
and
yxy + ay, + ++ + + Ont, = 1
Therefore
a= —da/2 j=1,2, mn
Thus,
¥ d
ms
or the parameter ) is given by
ra--2 and j= —%
. i
The corresponding value of a1? + 22 + +++ + aq? is
Note that the dual problem is to maximize the expression )\ aja; if
mA
> a; =1, and it is recommended that the reader solve this as an
a
exercise.
Example 2: Find the greatest and least values of z on the ellipse
formed by the intersection of the plane « + y +z = 1 and the ellipsoid
16x? + 4y? + 2 = 16.
Solution: Form*
F(xyyyzsAa,a) = 2 + aw + y + 2 — 1) + Aa(16z? + 4y? + 2? — 16)
Set the partial derivatives equal to zero:
F,=1+%+ 2d =0 @
Fy = + 8y =0 7)
F, =i + 82dr = 0 Gai)
and
atytz-1=0 Gv)
16x? + dy? + 2? — 16 =0 )
1 Since f(2,y,2) = z is to be extremized subject to the two constraints gi(z,y,2) =
aty+z2—1=Oand gx(z,yz) = 162? + dy? + 2? — 16 = 0.Instead of solving for the five unknowns, we note that (ii) and (iii) are
consistent if y = 4z[1 = d2 = 0 is excluded, since (i) is violated]. Sub-
stitution of y = 4z into (iv) yields x = (1 — 2)/5. Then (v) becomes
(in terms of 2)
16G —2)*, (416) —2)? | sige
ee
or, simplifying, this reduces to
Ql? — 322 — 64 = 0
The maximum and minimum values of z are therefore § and —§,
respectively.TWO
Introductory Problems
of the Variational Calculus
It is the initial objective of this chapter to introduce the concept of the
functional and then to extend the ideas of stationary values from functions
to functionals. The notion of the vanishing of the first variation of a
functional according to the theorem of minimum potential energy will
then lead to the Euler equations and natural boundary conditions. ‘This
will be done without stressing rigor in order to give the reader a feeling for
the scope of the variational method. (These ideas will be presented
more precisely in Chap. 3.) Examples of the static deflection of a beam
and a circular plate are developed to show how the differential equations
of equilibrium and boundary conditions may be derived. This is done
without recourse to free-body diagrams, which can often yield incorrect
field equations and boundary conditions in, for example, the case of
structural elements such as plates and shells. Finally, the chapter will
close with a brief discussion of some of the classical problems of the
variational calculus.
1 Concept of Functionals
Chapter 1 dealt with the subject of stationary values and more par-
ticularly with extrema, i.e., maximum and minimum for functions of oneor more variables. Similarly, the calculus of variations originated from
the endeavor to determine extrema or stationary values for functionals.
A functional is defined as a quantity or function which depends upon the
entire course or path of one or more functions rather than on a number of
discrete variables. The domain of a functional is a set or collection of
admissible functions which belongs to a function space or class rather than
to a region in coordinate space. Several examples should help to make
this concept clear.
Example 1: The length L of a curve y = y(z) defined over the closed
interval zo < x < 2, or more compactly [zo,2)], is given by the integral
L@) = [2 VIF UP a &
a (1.1)
Thus the value of L depends on the path or form of the function y(c),
called the argument function. In order that the integral (1.1) exist, it is
sufficient that y(v) be taken as an arbitrary continuous function with a
piecewise continuous derivative. This defines the space or domain of
admissible functions from which the argument functions may be selected.
Thus symbolically L = L(y) expresses this dependence of L on y.
Example 2: As a second example, consider the functional
$9) = [> “Fr @ at (1.2)
where s is a parameter independent of ¢ and thus is constant during the
integration. The “result” function f(s) depends upon the “object”
function F(#). We express this functional correspondence in the form
J(s) = L{FO} (1.3)
and f(s) is called the Laplace transform of the function F(t). The admis-
sible functions F({) are taken to be piecewise continuous in any finite
interval and to be of exponential order as i—> © ; that is, a positive number
M and a real number « exist such that |F(t)| < Me for all ¢ sufficiently
large. These conditions are sufficient but not necessary to guarantee
that f(s) exists for s > a; that is, the integral (1.2) converges.
Example 3: Functionals may also depend upon the behavior of more
than one function. In this example, we restrict ourselves to two admis-
sible functions y(x) and z(z), which are defined by the conditions
1. y(e) and z(2) are continuously differentiable on the closed segment
[20,21]
1 See, for example, R. V. Churchill, “Modern Operational Mathematics in Engineer-
ing,” Ist ed., p. 5, McGraw-Hill Book Company, New York, 1944.20 Two
2. y(x) and 2(z) have specified values at xo and 21; that is
y(t) = yo (ms) =
oe) = 2% ea) = a4)
A functional of y and z is
T(ya) = [2 Feyany's’) de (1.5)
where F is a continuous funetion of its five arguments x, y, 2, y',2".
Our considerations can be easily extended to accommodate n functions
of m variables involving, if m > 1, multiple integrals. For example,
Chap. 3, Sec. 2, pages 37 to 41, discusses the case of a single function of
two variables (n = 1, m = 2) in connection with making a double integral
stationary, where the class of admissible functions F is more severely
restricted than the integrand in (1.5).
2 Stationary Values for Functionals
The application of minimum-energy principles involves the definition
of stationary values for functions, or more generally, functionals, A
stationary condition is necessary in order that an extremum (maximum or
minimum) exist. In the case of a function of n independent: variables
S(e1,22, . . . ,tn), the conditions for a stationary value at a point P are
ef) _ of eae
Oxy |p ~ Axe |p cone ae ey)
as discussed in Chap. 1. In the case of functionals the problem is more
complex. For example, suppose that we are given the functional
Vw) = [2 Feyy'y") ae (2.2)
where F is a given explicit function of a, y, y’, y’’ and xo, a1 are fixed
numbers (Fig. 2.1). We seek the function y = y(x) (assuming its exist-
y
Fig. 2.1 Variation of extremizing
function with fixed ends.
yx) + By(x)
—— Ora
yx)
EXTREMIZING FUNCTION
(x05 Yo)ence) which makes V stationary where y and y’ are prescribed at z = xo
and z =. Let us examine the effect on V of a variation in y by a
small amount éy, where dy = dy’ = 0 at z — xo and x = a (Fig. 2.1).
If V is an extremized functional, then
v= cle y+ a + 3 oy" | ae = (2.3)
The left-hand side is called the first variation of the integral V.
In order to eliminate the variations éy/ and éy”’ from Eq. (2.3), the
second and third terms are integrated by parts:
okie: OF. Js pu d (aF
fo Sy BY de = [# af aia #) by de
2 OP an ay a fF ay (4 FY) gy]
5p "= [Soo -[e ar) of. co
2 da? (OF
ds (spr) ee
so that
aye ie [# d oF @ oF
jy aay tae xl by dz
ard oF ar
+[(2-z%) af + [Bw f-0 eo
If the boundary conditions (by = dy’ = 0 at 2 =z and 2 = m) are
imposed, then
nfor d oF, dt aF ic
av = [e-2% Ew vee -0 (2.6)
‘The above integral must vanish for all admissible 6y, which requires that
the expression in the bracket of Eq. (2.6) be zero, i.e.,
OF d (aF d@? (oF
oy ~ & (i) + (Sp) -° at
This follows from the fundamental lemma of the variational calculus
discussed in Chap. 3. The differential equation (2.7) is known as the
Euler differential equation corresponding to the functional defined by (2.2).
This, together with the boundary conditions, determines y(z).
Suppose that the requirements that y and y’ be prescribed at the ends
are relaxed. The competing curves would then appear as in Fig. 2.2.
For 8V to be zero, it is sufficient that (2.7) be satisfied in addition to
OF d oF oF
ay aay 9 Md By = 9 Gea
at x =a and x=. These equations (2.8) are referred to as the
“natural boundary conditions,” since they occur naturally from the22 Two
Fig. 2.2 Variation of extremizing
function with free ends.
-y(x) + By(x)
coe xex,
y(x) a
f EXTREMIZING FUNCTION
Ke Ko
variational process. Their physical significance will be discussed below
in the examples of a cantilever beam and a circular plate.
In general, the quantity 6V = 0 if at zo and 2; either
oF d oF
eee
OF
ay ~°
(@_ y is specified or =0
(2.9)
(ii) is prescribed or
and the differential equation (2.7) is satisfied.
THEOREM: We now state the following important theorems for
elastic bodies: the theorem of minimum potential energy and its converse.
Of all compatible displacements for a stable structure satisfying given
displacement boundary conditions, those which satisfy the equilibrium
equations make the potential energy V a minimum. (The potential
energy of a structure is the strain energy plus gains in potential energy of
the external forces.) The converse theorem, which is the one used most
often, is as follows: Of all the displacements satisfying the boundary
conditions, those which make the potential energy a minimum satisfy the
equilibrium equations. Since the potential energy is a minimum for the
true equilibrium state, it must also be stationary; ie.,
sv =0 (2.10)
Example 1: A uniform clastic cantilever beam (Fig. 2.3) with bending
rigidity EI and length L is uniformly loaded. Derive the differential
equation and boundary conditions by a variational procedure. Deter-
mine the end deflection y(L).2 A . EL is
Solution: "The strain energy is =" |," y’” dz (due to beam bending).
L
The gain in potential energy of the external load q is — f qy dx. There-
fore the total potential energy of the system is
which implies that
[EL ,
v=f [Fy a |de
Comparison with the general form (2.2) yields
EL
FaSu?—q (2.11)
To find the Euler equation (2.7) for this specific F we calculate
oF oF oF 7
gy Tt Gy tO gy Bly (2.12)
Substitution into (2.7) yields the differential equation of equilibrium
Ely" —q=0 (2.18)
The boundary conditions at the clamped end are y(0) = y’(0) =0
(deflection and rotation prevented).
The unsupported ends are free to deflect and rotate, so that the
boundary conditions (2.8) apply. Substitution of (2.12) into (2.8) yields
Ely" =0 (2.14a)
and
Ely” =0 (2 .14b)
Equation (2.14a) states that the shear is zero, while (2.146) implies that
the bending moment is zero at the free end. These are the natural
boundary conditions.
Fig. 2.3 Cantilever beam under
uniform load.24 Two
The solution of (2.13) may be effected by successive integration:
wn
qh
a
EI
wn He aoe
Oe ay +e, ¢, = arbitrary const
0 thus ¢, = — 7
BI
wn Vg
y Br e-)
na (e i
uw =ay(F- Leta) ee = arbitrary const
L
y"|,_, =0 therefore cs = 5
or
ma ge 2
u! = gay @ — We + 1)
Integrating again,
Buty’ |,_, = 0 and soe: =0. Finally,
gh (ee tae
if an(i ooo
where y i = 0 and consequently ci = 0. Thus the deflection of the
beam is given by
a rere
v= (S- $+ 2.15)
In particular, the end deflection is
= 7 (M_l IN _ dt
Yea ~ DET G 3 +o) 7 eT (2.16)
Example 2: A uniform elastic circular plate of radius b with a concen-
tric unsupported hole of radius ais shown in Fig. 2.4. The plate is loaded
axisymmetrically with a distributed normal force per unit surface area
q = q(r) at the top and is reacted at its outer edge as shown in the figure.
‘The differential equation and boundary conditions are required.ix
iG pelt SULLA
ke
Q
:
Fig. 2.4 Section of oe plate
with concentric hole loaded, and
reacted axisymmetrically.
Solution: The potential energy of the system (loads + reactions) may
be shown to be
2
V=aaD ds [rent +E+ anv ar — on f? qrw dr
+ 2n{rMw, — Qu)» (2.17)
where D = plate flexural rigidity
w = normal deflection (positive downward)
wy = dw/dr
Wr = dw/t ‘dr?
M = radial bending moment per unit of circumferential length
Q = radial shear per unit of circumferential length (The positive
direction is as shown in Fig. 2.4.)
= Poisson’s ratio
For stationary potential energy, 8V = 0 must hold. Noting that
20 Wer = 2
and that 54 = 4
8V = 2D i [2rd + 22 +=
w,*
d
a 6,
2 by, rf (2u, in) |
— nr in gr bw dr + 2n(rM bw, — 7Q 600)»
= 2D { rh ‘ [re Btbpp + (rte + 9) bu, + rw, 7 ar
6 gr bw _ 7Q bw
— PoP a+ (Gow D )}
Integration by parts yields
3
o- [ ew. + 1.) boy + (vm + 2) ww].
i ifs [z (rim + un) buy + 4 § (mm +2 on ar
- [oe +5 bu, 28) =0
lo Bs
1 This permutation of the order of operations is justified in Chap.
u
°26 Two
or integrating the term containing 6w, by parts, and collecting terms,
we have
ie [# (rtm + van) — 4 (rn + 9) - al bw dr
+ {(r%0 + ny) B06 + [~- $F Grog + ve) | on!
+(5 M sy, — a ow) =0 (218)
The differential equation is found by setting equal to zero the coefficient
of 6w in the integral with respect to r.!_ Thus
@ d w) gr _
oe
Since the terms in » cancel, we have '
fm~E(2)-§-«
or, expanding
ee
TWeere + Were — zo and y: > yo
by asmooth curve y = u(z) for which the time taken by a particle sliding
without friction from A to B along the curve under gravity is as short as
possible (Fig. 2.5). Let vo = initial velocity; then from energy considera-
tions, the velocity at any point (z,y) is determined by
v? — vo?
= gy — yo)
or
v? = Qgy + vo? — 2gyo
Let a be defined by
2gec = 2Wgyo — vo®Therefore v? = 2g(y — a), and taking the nonnegative square root of
both sides yields
a
v= FG = VM Vy
‘The time taken in the fall is
Pa eds c EVO
nD WS 2 ae
Ignoring the unimportant factor +/2g, the problem of the brachistochrone
may be stated as follows: Among all continuously differentiable functions
y(z) satisfying y(t0) = yo and y(z1) = y1, find the one which minimizes
r= [7 ew VIF OY
Vy —a
(3.3)
The solution to this problem was independently published by the two
Bernoulli brothers, James (1654-1705) and John (1667-1748). The
techniques at arriving at the solution were quite different. It is inter-
esting that a rivalry had arisen between these two men, and this gave an
unusual impetus and zest to the beginnings of the variational calculus.
Solutions to this problem were also obtained by Newton (1642-1727),
Leibniz (1646-1716), and L’ Hospital (1661-1704). More important, the
two papers of James Bernoulli in 1697 and 1701 were the starting point
for the researches of Euler (1707-1783), who was a pupil of John Bernoulli.
Buler was responsible for the first important result in the modern theory
of the calculus of variations.
Finally, we would like to mention briefly several other problems.
1. The simplest of all is that of finding the shortest smooth plane curve
joining two points A(zoyo) and B(z1,y:). ‘The length is given by
T= [?ViF Gide (3.4)
and J is to be minimized. Our intuition suggests that the answer is a
straight line connecting the two points. This is proved mathematically
in Chap. 8, Secs. 1 and 3.
In three dimensions, the corresponding problem is to determine two
funetions y = y(2), 2 = 2(x) such that the integral
JE VIF OFF CP ae (3.5)
has the least possible value, where y(z) and 2(z) are prescribed at the end
points of the interval.
Still more generally, we have the problem of finding the geodesics on a
given surface G(z,y,z) = 0, that is, of joining two points on the surface30 Two
Axo, ¥0)
Bix yi)
Fig. 2.6 Minimum surface of revo-
lution.
with coordinates (xo,yo,20) and (z,yx,21) by the shortest curve possible
lying in the surface. This problem is not simple.
Analytically, we seek among all triads of functions 2(1), y(0), 2(t) which
satisfy identically the equation of the surface
G(x,y2) = 0
and for which x(t), y(t), 2(4i) ( = 0, 1) are prescribed, to find that triad
for which the integral [" V+ 9 + dt is least.
2. The problem of minimum surface of revolution: Two points A (x0,y0)
and B(t:,y:), where 21 > zo, yo > 0, y1 > 0, are to be joined by a curve
y = u(z) lying above the z axis in such a way that the area of the surface
of revolution formed by rotating this curve about the z axis is least
(Fig. 2.6). The functional to be minimized is
T= 28 ["yVIF Gear cD)
This and some related questions have been approached both analytically
and experimentally using soap solutions.’
The problems introduced in this section are more extensively treated
in Chap. 3.
1 The reader is referred to a very informative discussion of the soap film experiments
in R. Courant and H. Robbins, “What Is Mathematics?,” 4th ed., pp. 385-397,
Oxford University Press, New York, 1947.THREE
Euler-Lagrange Development
with Applications
This chapter begins by returning to the basic problem of formulating
Euler's equation for a wide class of problems where the integrand in the
functional is of the form F(z,y,y’). The problem of the determination of
the shortest distance between two points is solved from both a variational
and a nonvariational viewpoint. The theoretical development is then
extended to more general integrands involving (1) more than one depend-
ent variable, (2) higher derivatives, and (3) several independent variables.
This is followed by a section on natural boundary conditions with applica-
tions to the problems of determining the minimum distance between
parallel lines and the minimum surface of revolution.
Still more general movable boundary problems with associated trans-
versality conditions are considered somewhat briefly. The brachisto-
chrone problem is then presented next and application to isochronous
clocks is considered in some detail. A systematic treatment of variational
notation follows; it shows the analogy between stationary problems for
functions and functionals. This notation is then applied to the small
deflections of a rotating shaft. The last three sections of this chapter are
devoted to Lagrange multipliers for functionals subject to the side or
constraint conditions in function and integral forms. Applications to
geodesic and isoperimetric problems are given to illustrate the method.
3132 Three
1. The Basic Problem—Euler’s Equation
Our object is to determine necessary conditions that a function y = u(z)
makes the integral
= [2 Feuy! de aa)
stationary where y(to) = yo and y(v:) = y: are preseribed. The function
F(z,y,y’) is assumed to possess continuous second derivatives with respect
to its three arguments. Also, y’’ is taken to be continuous in the interval
concerned.
To this end we construct y(z) = u(z) + en(x),! where (2) is contin-
uously differentiable and (0) = n(2:) = 0. The quantity ¢ is a param-
eter which is independent of 2. Substitution into (1.1) yields
T= 10 = [2F@utenu +e) a (1.2)
In order that J be stationary at e = 0 it is necessary that
a
F\ = 0 (13)
But
al a oF oF
mae (acta Fay * ow Fay *) i
and
al a [-[EerSeleene a
where F = F(z,u,w’). A a integration by parts of the second term
results in
2 OF oF ype (a aF
on Bul? = Gy? of a £ a & a) a
=- [*,(2 2
= Ox du!
since 7 is zero at each end of the interval. Therefore
dl n(@F _ d aF 2
mina (3-die) c.
However, 7 is arbitrary outside its differentiability requirements and
boundary conditions, Then, as shown in the proof of the fundamental
1 Note that in this case dy = en(z). Weshall be concerned with variations in which
both Sy and 6y' are small. These are referred to as weak variations.lemma of the variational calculus, (proved below) it follows that
au ~ daw ~° (8)
This is Fuler’s equation, and it is an ordinary differential equation of the
second order. It must be satisfied subject to the boundary conditions
u(t) = yo —U(as) = yn (1.6)
Proof of the Fundamental Lemma of the Variational Calculus
LEMMA: If a function ¢(x) which is continuous in [0,7] satisfies
the relation
[E 1@o@ ax = 0 (7)
where n(z) is any function such that
n(z0) = (zi) = 0 (1.8)
and 1/(z) is continuous in [x0,21], then
¢(z) =0 ins SrSu (1.9)
PROOF: Suppose on the contrary that $(z) #0 at x = £, where
zo <&0 in (ff:). Therefore hypothesis (1.7) is
violated and we have a contradiction. A similar contradiction is reached
if $(8) <0, which proves the lemma. In some applications the basic
lemma is required in a more restrictive form. If, for example, the integral34 Three
(1.7) is to vanish for all twice differentiable 7 for which (1.8) holds, then
by using the previous argument with n(x) = (x — &)%(% —2)* in
fo 0. Then since »/a? + y? > 2, it followsthat
[i vEFRaz fPea= frac =a
Therefore, L > a and the equality holds if and only if y = 0 or the curve
is the straight line. ‘This completes the proof and the example.
If the quantity d/dz F,, (in the general case) is determined by applica-
tion of the chain rule of differentiation, then (1.5) becomes in expanded
form (after multiplying through by —1)
wl Pu + U'Pan + Paz — Fu = 0 qa.)
where subscript denotes partial differentiation (e.g., Fvu = 0°F/du' du).
Every solution of (1.5) or (1.11) makes the given functional stationary.
The stationary character is necessary for the occurrence of a maximum or
minimum. However, as in the case of ordinary maxima and minima, it
is not sufficient to ensure the occurrence of either of these possibilities.
The determination of sufficient conditions is considerably more difficult
than the corresponding problem of the differential calculus. ‘The formal
systematic treatment of the sufficiency conditions as developed by Jacobi,
Weierstrass, and Hilbert, for example, is beyond the scope of the present
volume. The reader interested in this phase should consult references
by Akhiezer, Bliss, Bolza, Courant and Hilbert, Elsgole, Fox, Gelfand
and Fomin, and Pars.
2 Euler Equation for More General Cases
Integrals with More Than One Dependent Variable. The
problem of determining the stationary values of an integral can be
‘extended to the case of n dependent functions y; = y:(z),i = 1,2, ... ,m.
Let F(yrys, -- - Ym¥o¥» +» + sYn) be a function of the 2n +1
arguments z, yi, Y2,-.., y, which has continuous second deriva-
tives in the interval concerned. If we substitute specific functions
us = fila) = 1,2, . . . ,n) with continuous second derivatives into
T= [2 Fyn... evita) de (21)
then the result is a number. In general,
T= Tys,Y2, + + + Yn)
The variational problem now requires that we determine among all these
systems of continuously differentiable functions y:(x) (i = 1,2, . . . ,n)
prescribed at 2p and 23, the functions y; = u,(z) which make J stationary.
In many cases, the stationary value will be an absolute maximum or a
minimum.36 Three
We consider the family of functions
yz) = wr) ten) 1=1,2,...,0 (2.2)
where m(z), m2(z), . . - , m(z) are n arbitrarily chosen functions which
satisfy (xo) = 0, ni(t1) = 0 and furthermore that n;(z) possess continuous
first derivatives in [20,21]. Substitution of (2.2) into (2.1) results in
10) = [2 Futon... tebe ui pen, oo. su pent) de
A necessary condition for an extremum ys = 1(z) is that
leo = 9
But we ean choose, for example, mz = 95 = -* > = 1 =O and 1 #0,
so that.
160) = 2 FG, Ws $ ens, as us, «yy WL Henly hy 5 sh) de
Then, — as we did for the one-variable case of See. 1, there results
oF d a.
fl (je hiu) ne =0
which implies that
oF d oF
@
oe OF = Feats, «6. Unie, . 6. UA)
More generally, if u; alone is varied, then there results
oF _ d oF :
oo eS #=1,2,. n (2.3)
This is a system of n differential equations of the second order for the
n functions u(x), ..., Un(z). Each function u(x) must satisfy the
boundary conditions which are imposed on the corresponding functions
y(z)(@ = 1,2, . . . ,n).
Integrals Involving Higher Derivatives. The above development
may be extended to the case where the integrand assumes the form
Faeyy'y”, sy™). To be specific, attention will be confined to the
casen = 2. Therefore, we are interested in determining the requirements
on y(z) to make the integral
Ty) = [2 Fey y") de (2.4)
stationary subject to the conditions that y and y’ are prescribed at each
end.
Assume y = u(x) to be the function which makes J an extremum.Form the family of functions y(x) = u(x) + en(x), where ¢ is a constant
and 7 is an arbitrary function with continuous derivatives up to the fourth
order which together with its first derivative vanishes at the end points
of the interval.' Substitution into (2.4) yields
1 = [2 FG, ut ew! + of, ul + a1”) de (2.5)
Hall
Again
9 = 0 is required. We employ differentiation under the
integral sign using the chain rule of differentiation and then integrate by
parts to eliminate 1” and 7! using the boundary conditions on 7. ‘This
yields
foo[F Biot Bho ae = 0 26)
ze “Ga dete: :
where F = F(2,uju',u!’). Since n is arbitrary outside its differentiahility
requirements and boundary conditions, we deduce the Euler differential
equation corresponding to (2.4),
:
ret ret Sie =0 (2.7)
This is an ordinary differential equation of the fourth order. It must be
solved subject to the boundary conditions u and wu’ specified at x = zo
and z = 21.
Still more generally, if F(x,y,y/,y", - . . , y%) is the integrand and
y, y', y", . +4 yt) are prescribed at the ends of the interval,’ the
corresponding Euler equation is
da @ a oe in a i.) =
Pe — Pe + Per — Pert + (-I FF = 0
(2.8)
where F = F(z,uw'u, ... ,u®). This is a differential equation of
order 2n which must be satisfied subject to the 2n boundary conditions.
Integrals Involving Several Independent Variables. The above
method for determining the necessary conditions for an extreme value can
be applied equally well when the integral is a multiple integral. For
example, consider the instance of a double integral. Let D be a given
region bounded by a sectionally smooth curve C in the zy plane. The
integral to be treated is
I= J F(x,y,2,t02y0y) dx dy (2.9)
1 Euler's equations (2.7) and (2.8) still remain valid when the boundary conditions
are different from these in this section (see Sec. 4 of this chapter, on natural boundary
conditions).38 Three
where F is continuous and possesses two continuous derivatives with
respect to its five arguments and w(z,y) is prescribed on C. To effect the
extremization of (2.9) we introduce a one-parameter family of comparison
functions
w(ay) = u(ay) + en(z,y) (2.10)
where u(z,y) is the function which makes (2.9) stationary and n(z,y) has
continuous derivatives up to the second order and is zero on the bound-
ary C. The integral then becomes a function of e:
10 = [f F@yw,t0.s,) de dy (2.11)
4
which for stationary values must satisfy J’(0) = 0.
But
ye oF oF
10 = [f [Font feet $n ety
J :
and
r@= [ [ier inet oF 4, ac ay = 0
where nz = 41/dz, 1, = 4n/dy. To eliminate the terms in 7: and 7, we
integrate by parts—the former with respect to x and the last term with
respect toy. Therefore,!
od OF o OF
If o[ a sein ~ ayan |e”
OF dy aF dz], _
+ fo [ Ee |e =o (2.12)
where de = & ds, dy = ds, and dsis the differential arc length C. But
1 = 0 on C, so that the second integral is zero and then there results
OF a OF 0 OF
[fol ~ ia ou, ~ Hie] =
If we invoke an extension of the basic lemma of the variational calculus
(proved below), then the Euler partial differential equation of the second
1 The integral if means that C is followed in the conventionally positive sense of
description. This implies that the interior of the region D bounded by the closed
curve C'remains on the left as C’is traversed. More simply, (2.12) follows by applica-
tion of Green’s theorem in the plane.See os =0 (2.18)
is obtained.
Proof of the Extension of the Basic Lemma. Let D be a domain
of the zy plane for which
Jf emGtey) ax dy = 0 (214)
y
for all continuously differentiable (x,y) that are zero on the boundary C.
Furthermore, G(z,y) is continuous for all and yin D. We will establish
that G = 0 in D.
Suppose on the contrary that G # Oin D; then there exists a rectangular
subdomain D, of D for which G # 0. Let this rectangular subdomain be
i <2 < & m <1 m and assume (without loss in generality) that
G@ > Oinits interior. Choose 9(2,y) = (x — £:)?(@ — &2)*(y — m)*(y — m2)?
in D, and » = 0 over the rest of D. Then
[J reneew ae dy = [ n(ayy) Oey) de dy > 0
D !
contrary to hypothesis (2.14). Thus the assumption that G > 0 in Dy
has resulted in a contradiction. Analogously, we can show that G <0
in a subdomain also leads to a contradiction, which implies that G = 0 is
the only other possibility.
Example: Poisson and Laplace Equations—Static Membrane Deflec-
tion. The equations of Poisson and Laplace occur very frequently in
mathematical physies. Let us consider a problem which yields them
as Euler equations corresponding to making a particular functional
stationary.
Suppose that we wish to minimize the integral
ie {f fue? + uy? + 2f(z,y)ul de dy (2.15)
where f(z,y) is a prescribed continuous function and u(x,y) possesses
continuous first and second partial derivatives in D. Furthermore, u is
prescribed on the sectionally smooth boundary C of the two-dimensional
domain D, that is,
u=g(s) on€ (2.16)
‘The parameter s is the are length along C measured from a fixed (but
arbitrarily selected) point on C. Also, wis assumed to possess continuous
first partial derivatives on C.40 Three
For stationary J, it is necessary that (2.13) holds; i.e.,
Use + Uy =f (2.17)
which is Poisson’s equation. In particular if f = 0, (2.17) reduces to
Laplace's equation
Use + Uy = 0 (2.18)
For this example, we shall now show that the solution of (2.17) makes
the functional (2.15) an absolute minimum with respect to all functions
with continuous second partial derivatives which satisfy the boundary
restraints (2.16). Let (x,y) be any function of the set, and we write
U(z,y) = u@y) + @y)
where u(z,y) is the solution of (2.17) subject to (2.16). Since a(z,y)
also satisfies (2.16), the function n(z,y) = OonC. Define
AI = 1(@) — I(u)
fue + ma)* + (uy + my) + Fu + aM] de dy
5
4
- { [ut + ry? + 2ful de dy
= 2 ff tue + van Sade dy + ff tne + 98 ae dy
From Green’s theorem (or simply integration by parts)
[f (oon + wan ae dy = fon oe ag [J (1 + 1g) de dy
where d/dn stands for the derivative in the outward normal direction.
But then we have
a= 2 fongyas + [f 2 ~ (ee + uel de dy
+ ff Int + mA de dy
The first and second integrals are zero, since 7 = 0 on C and (2.17) holds.
Therefore,
at = ff int + nA1dz dy > 0
b
as the integrand is nonnegative. Moreover, AI = 0 if and only if
nz = ny = 0 in D, which implies that 7 is independent of both x and y or
merely a constant. However, » = 0 on C, so that 7 = 0 in D and on C.‘Thus AI > 0 if @ # u, which signifies that the function u minimizes the
functional I. This completes the proof.
Let us treat briefly a simple mechanical illustration which gives rise to
the integral (2.15). Consider the case of the static equilibrium of a
stretched homogeneous elastic membrane. By definition, a membrane
is an clastic skin which is incapable of resisting bending but does resist
stretching. In fact, the elastic membrane may be considered as the two-
dimensional analogue of the elastic string, which is treated in detail in
Chap. 6, Sec. 1, Example 1. It is assumed that the elastic membrane is
stretehed over a certain simply connected planar domain D bounded by a
rectifiable contour C. If u(z,y) denotes the displacement normal to the
plane, then the potential energy (for small deflections) may be shown to be
V=4 if] [Suet + uy?) — 2qul de dy (2.19)
The quantity S is the tension per unit of length. It is constant if it has
this constant value at all the points of boundary C. This means that
along any arbitrary cut through the membrane, the tension will be trans-
ferred from one side to the other without change of magnitude. Also
q = 4(z,y) is the external load per unit of projected area acting in the
direction of positive u. Comparison of (2.19) with (2.15) shows that the
Euler equation corresponding to (2.19) is
Use + Uy = — 3 (2.20)
and it is this Poisson differential equation of equilibrium which must be
solved subject to
u=0 on (2.21)
Solutions of the boundary-value problem defined by (2.20) and (2.21)
for particular domains such as rectangles or circles are readily obtained
by the method of separation of variables (in rectangular and polar
coordinates, respectively). The reader may refer to many sources for
detailed solutions; see, for example, R. Courant and D. Hilbert, pages
297 to 307. In this excellent reference book the free and forced vibra-
tions of such membranes are treated in an interesting and comprehensive
fashion,
Additional Examples
Example 1: Two Dependent Variables. Suppose that
T= J Cys — 2u? + (i)? — WD) de
is to be made stationary.42 Three
Then F(x,11,u2,uj,u3) = 2uyu, — Qu? + (ui)? — (uh)®, so that
oF
du, 7 2s — Ae
oe
Fa 720
The first Euler equation is [from (2.3)]
ard oF
San 7 de dat 7 22 ~ An — 2a’ =
or
uy — Quy — ui’ = 0 @
Similarly,
oF
uz
oF
aus
= Qu
= —2uh
so that the second Euler equation is
OF d oF
Su, de du,
or
um tu =0 Co)
Equations (i) and (ii) form a system of linear equations which is readily
solved. Differentiation of (i) twice and combination with (ji) and
thereby eliminating w, yields the ordinary differential equation of the
fourth order with constant coefficients
uy” + Quy’ + wr = 0 ii)
Let uw: = &* be substituted into (iii), which yields the characteristic
algebraic equation
O+)2=0 (iv)
with corresponding solutions } = +i, +i. This means that
us = (ext + ¢2) cose + (osx + cx) sine )
is the complete solution of (iii) where c1, cx, cy and cy are arbitrary (real)
constants of integration. Then substitution of (v) into (i) gives the
expression
us = ¢x(—2sin x + x cos x) + cy cost
+ (2 cosa -+2sinz) +esinz (vi)A simple check shows that (ji) is identically satisfied. The constants
¢1, C2, C3, Cs must be determined from the boundary conditions at each end
involving the quantities uw: and us.
Example 2: Integral Involving Higher Derivative. Find the Euler
differential equation associated with
JZ tou — W? + 6@)] ae
where ¢(z) is an arbitrary continuous function of x.
Solution: The integrand is
Fayy' iy") = ly? — (y)? + o(@)
where y’ is not explicitly present. Therefore
F(x,u,w ul’) = 16u? — (u’)? + $(2)
Fy = 32u Fu =O Fun = —2u"”
and the Euler equation (2.7) is
a a
Fy- Fw + ale =0
this yields the fourth-order ordinary differential equation
32u — uv” = 0
or
w” — 16u =0 @
The solution of (i) is readily found by substitution of u =e into (i).
This yields the characteristic algebraic equation
M—16=0
with roots \ = +2, +27. Therefore,
u = ce + cxe-** + cz cos 2c + sin 2x (ii)
where c1, ¢2, ¢s, and cs are arbitrary constants, is the complete solution.
Again, these constants are to be determined from the boundary restrainis.
Example 3: Find the Euler partial differential equation for the
functional
T= fff Peeyeni0, 0,0 ttee Dry es ay ele) de dy dz (2.92)
+
where V is a definite region in xyz space.
The functional (2.22) arises in problems of static equilibrium of elastic
solids and dynamics of “two-dimensional” bodies such as plates, shells,ccd Three
and membranes. In the former instance we have three spatial coordi-
nates, while in the latter, time replaces the third spatial coordinate.
We define the class of admissible functions w(z,y,z) as follows:
1. w and all its partial derivatives of order four are taken to be
continuous.
+ 2. w must satisfy on the surface S bounding V the linear forced con-
ditions of the form aw + 8 dw/dn = y, where a, B, y are prescribed
functions of the coordinates z, y, z.
Furthermore, F is assumed to possess continuous third partial deriva-
tives with respect to all of its arguments for all x, y, z in V.
To effect the extremization (again we assume its existence) let u(x,y,z)
make J stationary and write w(z,y,z) = u(z,y,z) + en(a,y,z), where 7 and
its first partial derivatives may be taken to be zero on S, since our imme-
diate interest is im the formulation of the Euler-Lagrange equation
corresponding to (2.22). Consequently, if we form J(e) and then
a
Flico = 0: there results
[ff toPe + nee, + mPa, + Po, + tesPese + tivP og + MP on
t e
t+ aya, + taFy,, + t2F u,,] dx dy de = 0
From the divergence theorem (or integration by parts), application of
the boundary conditions on 7, and the basic lemma of the variational
calculus, we conclude that
a a 3 a a
F.- -& fr) = a ey) -Z 2 +& 7 + i Poa)
2) + poe Pu,) + 525 Fun) = 0 (2.23)
= 5 z az in On
Equation (2.23) is the Euler partial differential equation. This must be
solved subject to the boundary restraints. Furthermore, we may derive
“natural boundary conditions” by not imposing all forced restraints on
w and u at the boundary. These conditions will occur by determining
that certain integrands of surface integrals must be zero. Natural bound-
ary conditions will be discussed in See. 4 of this chapter.
3 Solution of Euler’s Differential Equations
The Euler differential equation (1.5) of Sec. 1 is a second-order differ-
ential equation which in general cannot be explicitly integrated. How-
ever, in special cases the equation can be solved in quadratures.Dependent Variable u Absent—F = F(z,u'). For this problem,
F,, = 0, so that Euler’s differential equation becomes
J x) =0 (a)
which integrates to
Fy = ¢ = const (3.2)
Thus the problem of the determination of the extremization function is
reduced to a first-order differential equation. If Eq. (3.2) can be solved
for u’, such that
w! = Gee)
then
ua [Peeodth (3.3)
where k is another arbitrary constant of integration. In this instance,
the Euler equation is solved by quadrature. ‘The constants c and & are
determined from the conditions at the end points zo and 2 of the interval.
Suppose further that F = F(w), that is, both the independent and
dependent variables are absent. ‘Then
oF = 2 — sanction of w = (4)
which implies that u’ = k, where k is a constant which is a function of ¢,
that is, k = k(c). Consequently, the extremizing functions corre-
sponding to F = F(w’) are necessarily linear functions of x or straight
lines. Therefore, for the shortest distance in plane problem, treated in
Sec. 1, F = »/1 + (w)? and the solution is a straight line connecting the
two given points.
Example 1: It is required to extremize the functional
1) = [>MEOe 2>0
Solution: ‘The integrand F(w’z) = ~/1 + (w/z does not explici
depend onu. According to (3.2), the first integral of the Euler equation is
Pox = ¢ = const (3.6)
rV1 Fwy
Equation (3.6) is solved most simply by setting u’ = tant, —1/2 1 — wu? sin’ a 2 > 1 —sin Q = C8 >
Therefore the actual period T' is bounded by the estimates
cL an VL/g
Qn VE = aru =0. But
aH _ jn [30.mu 4 aoa"
Og dy OG | By’ dG
mS [ut av] ®
ag ie ;
={2 [Sea+ 28 ai]ae =0 j=1,2,...,N41
Integration by parts and utilization of 1j(¢0) = n(x) = 0 yields
nn [ae ae =
fon[%- & (3) |# =o f=1,2,...,N+1 @9)
where the 7; are arbitrary differentiable functions for which
1j(to) = 9;(@1) = 0
Thus
ae a9)
(3) - 0 (8.10)
which is a differential equation of order two. Integration of (8.10) will
give rise to two arbitrary constants of integration.
Thus we have N + 2 unknown quantities \1, X2, .. . , Av and the
two constants of integration. These may be determined from the N + 2
equations (8.2) and (8.3).
More generally, we now seck to extremize the integral
firey... 26m... ad oof 6.11)
with respect to continuously differentiable functions x(!), y(), . . . » 2(0)72 Three
subject to the N constraint conditions in integral form
Ky= [i Giaw-..att... ad 7 =1,2.-.50
(8.12)
The required equations to be solved are
a¢ _ d (a¢\ _
ad (2) ma
a¢ _ 4 (dd) _ 9
dy = dt \ay. (8.13)
oo a (eye
a di @) ae
where
n
o=F+ YG
ron
and F, Gi, . . . , Gy are defined by (8.11) and (8.12). The equations
(8.13) must be solved subject to the boundary and constraint con-
ditions (8.12).
Example: Find the closed plane curve of given length which encloses
a maximum area.
Solution: The curves which compete are taken to be of the form
z = x(t)
bo nates ars — Fe br dt = 0
mH
Integration by parts of the first term and use of the constraint condition
br; = 0 at t = t and tz yields
filer+ y s+ ars] dt = 0 (3.4)
a
where
T=} S mi? (3.5)
&
is the kinetic energy of the system of particles. Equation (3.4) is an
expression of Hamilton’s principle for a discrete system of particles in
general form (for a holonomic system). Various important specializations
can be made. First, the force acting on the ith particle can be divided
into external or applied forces F* and internal or constraint forces Fy,
so that
F, = Ft + Fe (3.6)
In many practical instances the work done by the constraint forces is
zero, i.e,
>; Fy: or, = 0 (3.7)
1
This is true, for example, for (1) the internal forces in a rigid body,
(2) the reaction on a body rolling (without slipping) on a fixed body, and
(8) the tensions in an inextensible string of an Atwood machine.
For this situation, Hamilton's principle reduces to
fi ler+ 5 Fe ori] dt = 0 (3.8)
4
Finally, if the applied forces are derivable from a potential function V,
we have
aV aV av
Fe=— @ x, #) (3.9)
where zi, ti2, aia are the rectangular coordinates of the ith particle.Alternatively,
(3.10)
and therefore for a conservative system,
a fs a
[" or — avy) a= sf" —v)a=5 ("La =0 (11)
In words, Hamilton's principle states that the motion is such that the integral
of the difference between the kinetic and potential energies of the system is
stationary for the true path when the initial and final configurations are
specified.
4 Lagrange’s Equations with Applications
The Lagrange equations may be deduced immediately from Hamilton’s
principle for a conservative holonomic system. From (3.3)
.yn
The kinetic energy of the system is
T= 4) mite = Tass, «+ + dordudes « « « sdnt
Ft
Similarly, from (3.3),
V = V(rurs, . . . sta) = Vgugn - - - gost)
so that the Lagrangian
L=T—V = L(gygs, - - - sdorduda, » - + Gost)
From condition (3.11) and the basic Euler formula there results the equa-
tions of motion
doaL_ aL
tae sp (4.1)
or equivalently
Golaor ev. ie
HG 3 7 age TTP (4.2)
‘These are the Lagrange equations for a conservative system. The
advantage of the Lagrangian equations is that the equations governing
the motion of the system (with ideal constraints) can be written down by84 Four
k
ma
™ www
Eo Tl
Fig. 4.2, Basic two-mass-spring sys-
tem performing translational vibra-
tions.
tule, thereby enabling the practitioner to find them with a minimum of
mental effort. On the negative side, this procedure does not promote
insight into the physical processes that are taking place. In short, the
use of Lagrange’s equations is analogous to the function served by integral
tables to the user of mathematics. While these tables do not give insight
into the integration process, they expedite the ensuing calculations and
considerably reduce the chance of error. The application of faulty (but
reasonable appearing) free-body diagrams for complex systems such as
plates and shells leads to the formulation of erroneous differential equa-
tions, thereby emphasizing the need for the Lagrange method.
Three examples will be demonstrated to illustrate the formulation of
the Lagrange’s equations for oscillating systems.
Example 1: Linear Oscillations of Two Coupled Masses. Consider
the particular case of small free vibrations of the system shown in Fig. 4.2.
‘Two masses m; and m: can slide without friction along the horizontal
zaxis. The connecting springs have spring constants given by ky and he.
We seek the equations of motion for the two-mass system.
Let x1 and zz denote the respective displacements of the two masses
from their equilibrium positions. The kinetic energy of the system is
T = 30m? + mets?)
while the potential energy (which in this case is the strain energy) is
given by
V = Heys? + Ha(er — 21)?
Lagrange’s equations are (L = T — V)
d (aL
dt \ oR,
so that we have
—kyer + heo(a, — 21)
—kn(e2 — a1)
midy
mtr
(4.3)vertical oscillations where the displacements are taken from the static
equilibrium position.
The above system is completely analogous to the torsional oscillations
of two circular disks suspended from two coaxial circular shafts as shown
in Fig. 4.3.
The equations to be solved are analogously
T
Gi, Gals at be
Th =~“ a+ F — 0) iL | ;
Gal:
The = —"F7 — 0) 1
T
where 63, #2 = twists of the disks i ir
Gi, Gz = shear moduli of the shafts [ ;
J;, J2 = axial polar moments of inertia |
of shafts cross sections
wast wd.* Fig. 4.3 Torsional oscil-
32° 32 lation of circular shafts.
Iy, Lz = lengths of the shafts
J;, Iz = axial moments of inertia of disks
and quantities G,J:/L; are the “spring constants” of the shafts.
Example 2: Find the differential equations governing the free
vibrations of the system in Fig. 4.4a (about its static equilibrium position)
77
Fig. .4 Vibrations about static
equilibrium.
eee fee
Bie a 1
:
na I 2q2-4)
Se
we
()86 Four
assuming that the springs are massless and that the bar to which masses
m are attached is rigid and massless.
Solution: Let q, and q: be the displacement of the masses (Fig. 4.4b).
The system has two degrees of freedom. We choose q: and gz as the
generalized coordinates. ‘The rigidity of the bar requires that the dis-
placement of the other spring is 292 — qi.
The kinetic energy of the system is
T = ¥(mix? + més’)
and the potential energy of the system (strain energy in springs) is
kg? ik
= +5 Cn — a)?
Lagrange’s equations become
mgr + 2k — m2) = 0
m2 + 2k(2q. — qi) = 0
so that the mode shapes and natural frequencies may be readily obtained.
Example 3: As a final example, we consider the determination of the
equations of motion of a spherical pendulum (Fig. 4.5). ‘The position of
the particle of mass m is completely defined by the spherical coordinates,
namely, the colatitude @ and the longitude @. Thus we choose these
angles as generalized coordinates: q: = @ and q: = ¢. Then the kinetic
Fig. 4.5 Spherical pendulum.and potential energies are
2
T= oe (6 + 6% sin? 8)
V = mgacos@ — (VJpazn = 0 is chosen as datum)
and differentiation yields
—= 2 2 2 gi:
0 ma? ma? ¢? sin @ cos 0
oF = mga sin 6
a6
or : or av
pelle ey ee a
36 ma? ¢ sin?? 6 0 6 0
Therefore (4.2) yield the following equations of motion for a spherical
pendulum:
ma*§ — ma? ¢? sin 6 cos @ = mga sin 6
d j
ay (ia? ¢ sin? 8) = 0
The particular cases of a simple pendulum (¢ = 0) and the conical
pendulum (6 = constant) may be readily treated.
8 Lagrange’s Equations for a General System
We seek the Lagrange’s equations for a general nonconservative
holonomic system of p degrees of freedom.
The inerement of work done 6W by the resultant force field F; during a
virtual displacement dr; of the ith particle is
oW = Y) Fie ors (6.1)
ron
But from r; = ri(qi,g2 . . . ,Qp,t) it follows that
>
ors
a= Y ag (5.2)
hh qe
Substitution of (5.2) into (5.1) yields
2s
= peed
aw = > Fis 5 ba (53)
ee ia
(Result of sum is independent of ordering.) Define a new quantity Q,88 Four
called the generalized force component by the formula
2
aW = Y) Qu du (6.4)
mo
so that
os
a= Sr. ar (5.5)
The generalized force Q; has the dimension of a force when q is a length
and the dimension of a moment when q: is an angular coordinate. In
general, the dimension of Q, is determined by the rule that the product
Qiqi has the dimension of work. It is not necessary to use (5.5) to caleu-
late the generalized forces, since they can be found directly from (5.1).
Consider, for example, the spherical pendulum. The only external
forces are the weight —mgk (where k is a unit vector in the positive z
direction) and the tension in the string. Then the virtual work corre-
sponding to a virtual “displacement” 66, 6¢ is
8W = —mgk - [a 80 ey + asin 8 6¢ es]
where ey and eg are unit vectors in the direction of increasing @ and 9,
respectively (the string tension does no work). Thus if @, ® are the
generalized forces in the @ and ¢ directions,
8W = © 0 + © 54 = mgasin 656
so that the generalized forces are
@ = mga sin 6
o=0 66)
Let us return to the problem of formulating the equations of Lagrange
for a general (conservative or nonconservative) holonomic system of
n particles with p degrees of freedom. Substitution of (5.4) and (5.1)
into (8.4) yields
fort 3 Q dqx) dt = 0 (0)
or if we perform the variation in the usual way (integrate by parts and use
the fact that gi is zero at time t = t; and ¢ = fy), there results
adoT
ik 3 (B- a8 +e) naoHowever, the quantities 64 can be varied independently, so that
det at
Hag ~ au 7 k=w1,2...,7 (5.8)
If the system is conservative,
av
a= - 5, (6.9)
and (5.8) reduces to (4.2). If some I < p of the generalized forces are
conservative while the remaining are not, then (5.8) may be replaced by
d eT _ aT , aV
He Gu tag 9 RELA. d a0
daT_ aT _ hale 6.10)
Hin ~ ag 7 =1+1,1+2,...,p
It is important to emphasize that (1) the generalized coordinates can be
chosen in a number of ways, but the form of the Lagrange equations
remains the same independently of the choice and (2) workless constraints
are automatically eliminated, since they do not contribute to the expres-
sion for 6W.
Consider next the particular case when the kinetic energy is a function
of the generalized coordinates q: and the generalized velocities 4 and not
explicitly of the time. This will occur when the displacement vectors are
functions of the generalized coordinates only, i.e.
12
Also, the potential energy V is assumed to depend only on the generalized
coordinates and not explicitly on the time.
Thus the integrand 7 — V in Hamilton’s integral does not depend
explicitly on the independent variable ¢. It is therefore permissible to
use the first integral of Euler’s equation:
M92)» + Gp)
2
T-V— > & rT = const (6.11)
mh
which may be rewritten in the form
-
—TH+ V4) Got =H = const (6.12)
he ee
But 7 is a homogeneous quadratic function of the form
a)
Dy
disdids (5.13)
jar
i90 Four
where
by = by (6.14)
Then
Cer yoet) recy
an = Del he + 3g, %
But
ag _ {1 ifh=j
Se (0 fk 5
since the generalized velocity components are independent. Con-
sequently,
2 2
= d, bade +) bess
ry
or from (5.14),
ee
oa? > bade (5.15)
so that
2 ae
. oT a
” 55 7 2 2 Dade = 27 (5.16)
Substitution into (5.12) yields
T+ V =H = const (5.17)
Therefore the principle of conservation of energy has been deduced, namely,
that during the motion the sum of the kinetic energy and potential energy does
not vary with the time.
6 Normal Oscillations
This section will give a more detailed analysis of the free vibrations of a
conservative system about a position of stable equilibrium. Consider a
dynamical system with a finite number of degrees of freedom. It can be
shown that the most general motion of such a system which is disturbed
from a position of stable equilibrium may be represented as a superposition
of independent periodic motions known as principal or normal modes.
The number of such special modes of motion is equal to the number ofdegrees of freedom. These particular motions are characterized by the
following properties
1. The motion of every particle of the system is simple harmonic.
2. The period and phase of the simple harmonic motion are the same
for all the particles.
3. The ratio of the amplitudes of the displacements of the normal
modes is determined by solving a homogeneous linear system of equations.
4, The normal periods (and frequencies) are found by solving a deter-
minantal equation or by iterative matrix techniques.
To illustrate the ideas, let us consider the vibrations of a light elastic
string of length L; + L: + La stretched between the two points A and B
(Fig. 4.6). Let two particles of mass m, and m, be located at the distances
L, and L; from the ends. The particles are assumed to be supported on a
smooth horizontal plane, or, equivalently, gravity can be neglected.
Initially the particles are at rest and the tension & in the string is
constant throughout. The particles are given a small displacement
perpendicular to the rest position and then released. It is required to
find the resulting oscillations.
‘The configuration at any time t (and somewhat exaggerated) is shown
in Fig. 4.6.
‘The changes in length of the three string elements are
2
Als = Lise 1) = 48 ga, 2,8
Since the potential energy V is the tension k multiplied by the increase
in length and
oe 4e8Te md eB
we have
—klg?, Ga)? , a
Vea3 [% i iy
The kinetic energy is T = }(mgi? + mags®), and the Lagrange equa-
Fig. 4.6 Vibrations of a light elastic
string with two attached masses.92 Four
tions (4.2) are
mas tk (P+ OH 7) =0
(1)
mage +k (52 a +E e =0
In order to simplify the formulas (without masking the concepts) we
proceed to the special case m, = m: = m; Ly = Lz = Li = L.
Lagrange’s equations become
mas + Ms a Ha. — %) 9
(6.2)
mi + _ a) , ka _ 9
On examining these equations, it is noted that, when q: = 2 or q1 = —q»,
the two equations become identical, namely
mg + fq =0 (6.3)
when qi = gz, and
mg + & q=0 (6.4)
ifq = —g.
These equations are expressions of harmonic motion and may be
interpreted as follows:
1. If the two masses are started with equal amplitudes on the same side
(q = q), the motion will be harmonic with frequency w: = /k/mL.
The center string will then remain horizontal, and the masses, being
unable to exert a vertical force, cannot influence each other (Fig. 4.7).
2. If, on the other hand, the two masses are started with equal and
opposite amplitudes g: = —q2, the motion is again harmonic with fre-
quency «2 = ~/3k/mL but the two masses will remain out of phase with
each other (Fig. 4.7). Note: These are the only two starting conditions
for which the motion is harmonic, as can be shown by examining the
original equations.
Let us return to the simultaneous equations (6.2) and a more direct
procedure for determining the normal modes of vibration. Introduce
the quantity
k
Meo (6.5)Fig. 4.7 Normal modes of vibrations.
so that equations (6.2) become
hi + Wo — Mg: = 0
di — Xn + 2N%Q2 = 0 6.6)
Let us try solutions of the form
qi = @ cos (nt + e)
2 = B cos (nt + 6) 6.7)
where a, 8, n, and e are constants, Substitution of (6.7) into (6.6) yields
the homogeneous linear system:
a(—n? + 2n) — Br? = 0
aan =o (6.8)
‘This system possesses a nontrivial solution if and only if the determinant
of the coefficients
nt — 2° oN
PNG n® — 2n?
or, by expansion,
nt — 4dr? + 34 = 0 (6.9)
which has the solutions
mt= mt = 3 (6.10)
For each 7%, the ralio of 6 to a is readily found. For n* = ni? = 24,
B/a = 1; and for n? = nz? = 30%, 6/a = —1. Thus
qi = = Cos (At + 41)
q: = a cos (At + e1) (6.1)
represents one set of normal modes. Similarly,
a 2 08 (4/3 At + €2) 6.12)
q2 = —a2 cos (4/3 M+ «)94 Four
is a second set of normal modes. The complete or general solution is
obtained by superposition, namely
qi = c4 Cos (ME + €1) + a cos (4/3 Mt + €)
gz = a1 cos (At + €1) — a2 cos (4/3 Mt + &)
This is the most general solution, because it contains four constants a1, a2,
4, €2 which may be adapted to satisfy arbitrary initial conditions corre-
sponding to given positions and velocities of the two particles at t = 0.
(6.13)FIVE
Deformable Bodies—
Theory of Elasticity
The first portion of this chapter will be devoted to the geometrical
concept of the relative deformations of a deformable body. The material
of the body under consideration is taken to be a continuum, i.e., the
atomistic structure of matter is disregarded. For example, when we
introduce the mass density of a body “at a point P,” this means the mass
per unit of volume at the point. In other words, if P is surrounded by a
portion of the body of mass Am and Volume AV, then the density p of
the body at P is defined by p = iim. SP? where it is assumed that the
limit exists and is a continuous funetion of position. ‘Therefore the body
is replaced by a continuous mathematical model whose geometrical points
are identified with the corresponding material points of the body. It is
important to note that the range of applicability of the concepts in the
sections on strain and the following Secs. 4 to 7 on stress applies equally
well to elastic! or even more general inelastic solid bodies. Furthermore,
these formulas also apply to nonsolid bodies such as liquids and gases.
Therefore, these concepts may be applied, for example, to problems in
elasticity, plasticity, aerodynamics, and hydrodynamics.
1A deformed body is called elastic if it possesses the property that it regains its
original unloaded shape when the forces causing its deformation are removed.
9596 Five
Starting with Sec. 8, our considerations will be restricted to linear
elastic bodies, i.e., those for which the stress and strain components are
linearly related. Examples of materials which essentially exhibit this
linear relation in considerable load-deflection ranges are high-strength
aluminum alloy, low-carbon steel, and glass. The logical extension of the
basic Hooke’s law, termed the generalized Hooke’s law, is treated in
Sec. 9. It is used in all developments of the linear theory of elasticity
and is applicable to the most general case of anisotropic elastic bodies. It
involves 36 constants, but with the assumption of a strain energy density
function the number of elastic constants reduces to 21. Furthermore, by
elastic symmetric properties the number of elastic constants may be
further reduced, and we have various crystal classes depending upon the
assumed levels of symmetry. Finally, when the elastic properties of the
body are the same in all directions, the body is elastically isotropic and the
number of fundamental elastic constants reduces to 2. The physical
significance of the isotropic elastic constants is treated in Sec. 10. Sec-
tion 11 is concerned with the intimate connection between the strain
energy density function and stress and strain components. The expres-
sion for the strain energy is derived for isotropic structural elements.
1 Strain
Consider an arbitrary point in a three-dimensional elastic body B
(Fig. 5.1) which before deformation has rectangular coordinates (z,y,2).
Let: (u,0,10) denote the displacement field of the point at any time t so that
u = u(a,y,2), » = v(e,y.2), and w = w(x,y,z). Therefore, after deforma-
tion the rectangular coordinates of the point are given by
vw =2+u(zyz)
y =y + vey.) (Lt)
z+ w(zy,2)
Q
ds Plictu,yty,z+w)
Possy2)
8 Fig. 5.1 Vector displacement field
for two neighboring points P and Q.Let (ds)* denote the square of the distance between two neighboring points
before deformation and (ds’)? denote the corresponding distance after
deformation. From (1.1),
dx! = de + du = dx + tz da + uy dy + us dz
dy! = dy + dv = dy + vez +0, dy +0. dz (1.2)
de! = dz + dw = dz + wz de + w, dy + w, de
so that
(ds')? = (da!)® + (dy')? + (de)?
= [dx + uz dz + uy dy + u, dz}?
+ [dy + v2 da + vy dy + v, dz]?
+ [dz + ws dx + w, dy + w, dz]?
or
(ds’)* = (dx)? + (dy)? + (dz)*
+ Yece da? + Ley dy? + ers de?
+ ery die dy + eye dy de + ees de dx
+ ese dx dz + Leys dy dz + 2exy dz dy (1.3)
where
Us + $(us® + v2? + wz")
vy +H (uy? + vy? + w,?)
ws + 3(us* + v2? + w,") (1.4)
ye = B[Uy + Ve + Ustly + vary + Wetdy]
$l. + wy + Uytls + oye + wywr]
Gz = ez = Ewe + Ue + Uets + 002 + wwe)
We introduce the quantity ¢ defined by
_ ds’ —ds
-2 = (1.5)
and called the strain in the line element originally of length ds. As repre-
sented by (1.5), ¢ (which is greater than —1) is the change in distance
between the points P and Q (due to the deformation field) divided by the
distance between them before deformation. A simple calculation shows
that
— (as? = (ds)?
dst
et+ie= (1.6)
so that if the initial direction cosines of the line element are (lym,n), then.
e+ de? = eral? + end + eadln
+ eal + eqn? + exmn
+ earl + eynm + en? (17)
Thus from (1.7), in order to calculate e it is sufficient to know the six98 Five
coefficients €22, €yy, €s; xy) €ys) €z, Which are expressed in terms of the
displacement components according to (1.4). In particular, for a rigid-
body displacement, the six quantities are zero. Note that subscripts
(x,y,z) connected with (u,v,w) denote partial derivatives (for example,
v2 = dv/an).
The symmetric matrix (czy = €yz, €ye = €sy Gz = €xz)
fez fay Exe
ei fy fe
Gz Gy Gee
is called the strain tensor, since its components obey the tensor law of
transformation of a Cartesian tensor of rank 2, as shown in the next
section.
The components és, éy, and ¢, are readily interpreted physically. If
we set 1 = 1, m = 0, n = 0 in (1.7) (corresponding to the case of a linear
element which before deformation is parallel to the « axis and is of length
ds = dz) then
Cz + de2? = tox (1.8)
If (1.8) is solved for ez, and noting that e, must be greater than —1, we
have uniquely
é: = V1 + 22 —1 (1.9)
For most engineering applications, e.g. where aluminum, structural steel,
etc. are employed, the components of strain are very small in comparison
with unity (¢.g., in many applications |e,| < 0.001), so that
fez = Ce (1.10)
Thus, the component e,, of the strain tensor represents the extension or
change in length per unit length of an element originally parallel to the
z axis. Similarly, the relative extension of a fiber or element initially
parallel to the y and 2 axes may be expressed analogously as
oo. (1.11)
6 = V1 Ft 26 — 1
Next we shall deduce an interpretation of the strain components
éy €m and éz. ‘To this end, let 0 and 0, be two neighboring points at a
small distance r apart. If J, m, n denote the direction cosines of the ele-
ment 00; and (6z,éy,6z) the rectangular coordinates of 0; with respect
to O, then
bx = rl
by = rm (1.12)
éz = rnIf (w,w) are the z, y, and z components of displacement of point 0, the
displacement of point 01 is (by Taylor's expansion)
ty = Ut Us bx + tty by + Ue Be
v1 =v + ve bx + dy by + ve bz (1.18)
wy = w+ wz bx + w, by + w, be
where terms of second and higher order in 6z, dy, and 6z are neglected
because the quantities 6x, éy, and éz are numerically very small in com-
parison with unity. Owing to the deformation, let 0 move to O while
Ox is displaced to O;. If r(1 + e) is the distance between the two points
after deformation and if (l1,m1,n1) are the direction cosines of segment
joining 0; and Oy, then, using (1.13),
y= etme _ At us) te + wy by + te be
Oe Eile Oe
7 + €) 7d +6)
_ dy tur —o _ vede t+ (1 +0) by + &
ere a ee aa
_ be + wi —w_ vedo + wy dy ++) &
cen at)
Substitution of (1.14) into (1.12) yields
(+e) = (+ us)l + uym + un
(1 + e)mi = vel + (1 + 0,)m + vn (1.15)
(1 + e)m: = wil + wym + (1 + w,)n
Similarly, if we consider a second element of length 7 through the
same point with direction cosines (I/,m’,n’), there results
(+e) = (+ wa)! + uym! + un
(Lt e’)mi, = vd! + (1 + 0,)m! + on’ (1.16)
(tent = wal! + wym! + (1 t+ w,)n’
If each equation in (1.15) is multiplied by the corresponding equation in
(1.16) and addition is performed, then, utilizing the definitions (1.4),
(1+ e)(L +e!) + mami, + many) = Ul + mm! + mn’
+ Qlezall! + eymm! + enn!
+ xy(Im! + I'm) + eye(mn! + m’n)
+ ee(nl’ + In’)| (1.17)
In particular, if the two elements of lengths r and r’ are orthogonal, then
W' + mm! + nn’ = 0, and denoting @ as the angle between the elements
after deformation (so that cos @ = lil; + mym; + ninj), we have
G+ 6)0 +e’) cos 6 _
2
eeall! + eymm! + enn! + en(lm’ + Um)
+ ee(mn! + min) + ee(nl! + n'l) (1.18)100 Five
Specializing still further, we choose
l=1,m=0,n=0 and Jl =0,m' =1,n'=0
so that the originally orthogonal segments are parallel to the x and y
directions, respectively. Substitution into (1.18) gives
A+ OG +e) e080 19)
The angle ¥ = 5 — 0 is known as the shearing strain. If we note that
the three quantities — 0, ¢, and e’ are generally very small in magnitude
2
in comparison with 1, then
Y= eay (1.20)
Again, the other quantities ¢,, and «2 may be interpreted analogously as
approximately one-half the shearing strain of segments originally parallel
to the y and z axes and the z and z axes, respectively.
2 Transformation of Strain Components—Principal Axes
Next we seek the transformation of components of strain at a point
from one Cartesian coordinate system to another. Let (l,m,n) and
(Um! ,n!) be the direction cosines of an element with respect to the two
systems, namely, (z,y,z) and (z’,y’,2’). Also, we define the direction
cosines by the following table:
2 y z
a 4 my, m
y ly ms Mm
2 L m na
2
where, for example, 1s = cos (z,2'). From (1.7), since ¢ + 5 is inde-
pendent of the particular Cartesian system,
eed? + eqn? + ean? + eal + 2eumn + eanl = eva (U!)?
+ eval’)? + ere(n!)? + eryl’m! + 2eyym'n! + ern! (2.1)
But
L= bl + bam! + lin’
m = ml! + mam! + msn! (2.2)
n= nal! + nym! +n!
Substituting (2.2) into the left member of (2.1) and noting that I’, m’, and‘n’ are arbitrary within the constraint condition (’)? + (m’)? + (n’)? = 1,
there results
eet = Ly%€ee + matey + tatece + Qlimr€sy + 2mamreye + Balers
eyy = Unters + matey + natece + Lamatay + Qmanaeye + 2aleee
eve = [sez + matey + mae + Qamseey + mame: + WMalseee
cory = Lalates + mitmatyy + Matacer + (Lama + Lams)ery (23)
+ (mana + mumi)eys + (mile + Melrose
ates + MeMaeyy + Nase + (lems + lam2)exy
+ (mans + mana)eys + (Mala + Nsla)ere
eer = Uslites + mamrey + Natiers + (Lym + lims)exy
+ (mani + mins)eye + (mali + mils)ece
It may be recognized that the strain components are Cartesian tensors
of rank two because of the form of the transformation. This becomes
more apparent when indicial notation is used.
From the expressions for the transformation of the strain components
it is a simple matter to determine the direction for which the direct com-
ponents of strain are extremized. These are the directions of principal
strains. From (1.9) it is clear that maximizing ¢. is equivalent to
maximizing the strain or relative deformation e.. Let 2’ denote the
required direction which extremizes evs as given by the first formula of
(2.3). The direction cosines lh, m1, m1 are connected by the equation
L? +m? +n2—1=0 (2.4)
Thus the Lagrange method of undetermined multipliers may be used.
To this end we form the function
eye
A = exp — ely? + mye +1? — 1) (2.5)
where e is the Lagrangian constant multiplier, and then set
oH _ oH _ ay Lh
Performance of this differentiation results in
(cz — Oli + exymi + exxtrr = 0
eed + (Ew — 6a + thy = 0 (2.7)
€zal1 + €y2m1 + (Ce — €)m1 = 0
The system (2.7) has the trivial solution], = m, = ni = 0, which must be
rejected because it contradicts the constraint condition (2.4). For non-
trivial solutions, it is necessary and sufficient that the determinant of its
1 Detailed treatment of indicial notation and the elements of Cartesian tensors is
given in Appendix A. For further applications of indicial notation, the reader is
referred to Chap. 6, starting with Sec. 3.102 Five
coefficients be zero:
tes € bay
fy fw 0 (2.8)
fae ye
Expansion of the determinant yields the following cubic equation for the
determination of the principal strains
é-he+he-Ih=0 (2.9)
where
Ty = eee + Gy + Ge
Ta = tasty + €yyts + €ss€ez — Exy?
= ey? — Ges? (2.10)
fez xy xs
Ts= ley Gv &
cs Gye as
Equation (2.9) is a cubic equation and therefore must possess at least one
real root. Furthermore, because of the symmetry of the determinant
(2.7), it is known that all roots must be real. Denote these roots as , €2,
and ¢. Since the values of the principal strains must be independent of
the particular Cartesian coordinate system, the coefficients I;, Is, and Is
which enter into (2.9) must be invariant quantities, ie, have the same
values for any rotated coordinate system. The quantities 1, Is, and I;
are designated as the first, second, and third invariants of the strain
tensor.
Associated with each principal strain there is a definite direction found
from the solution of system (2.7) (which determines the ratios of the direc-
tion cosines Js, m1, m1) and the constraint condition ly? + m1? + mi? = 1.
It will now be established that the three principal directions are mutually
orthogonal. Let ¢1, ¢2, es be the principal strains and (11,m1,n1), (Inman),
and (ls,ms,ns) the direction cosines of their respective directions. Then
from (2.7), corresponding to ¢ and e2:
ely = xaly + cay + eta (2.11)
ey = eal + ey + et (2.110)
ext = zal + eet + eat (2.110)
and
eale = eeala + enya + €xxtta (2.12a)
exttty = exyla + equa + eta (2.126)
eat, = ela + ea + esta (2.12¢)
If both sides of (2.11) are multiplied by Js, (2.118) by ma, and (2.11¢)by nz and the resulting equations are summed, we have
eallale + mame + mina] = eselile + eyymime + examine
+ exy(male + mals) + eye(nm2 + nym) + ee(lime + lam) (2.182)
Similarly, multiplication of (2.12a) by l, (2.12b) by ms, end (2.12c) by m
and adding yields
eq{lile + myms + nine] = exalila + eyymyme + een in2
+ exy(tmale + mals) + eye(name + mam) + ece(lime + lena) (2.180)
Subtracting (2.13b) from (2.13a), there results
(4: — €)[lile + mym2 + nine] = 0 (2.14)
In general «: * es, which implies that
le + mymz + rum = 0 (2.15)
and proves the condition of orthogonality. (If 1 = ¢2, then all directions
in the plane of the two principal axes can be shown to be principal
directions and orthogonality may still be satisfied by selecting directions
which are perpendicular.) Ina similar manner, lils + mms + mits = 0
is readily established, which proves that the directions of the three prin-
cipal strains are mutually perpendicular to each other.
It can also be shown that if the principal strains are such that
© > @ > e, then «: is the absolute maximum direct strain and ¢; the
minimum, while ¢ is neither a maximum or minimum.
Suppose now that the 2’, y’, and 2’ axes correspond to principal direc-
tions, so that eve, ¢vy, and éve are principal direct strains. We shall now
establish that the corresponding shearing strains ey, «', e’ ate 7eT0.
From (2.3) we have
eey = lallress + Mr€sy + Mees] + mallreay + Mey + Maeve]
+ malliere + mitye + Miés] (2.16)
Replace the expressions in brackets by their equivalent expressions
according to (2.7):
Lafels] + mafems] + nafena}
e{lils + mim2 + mina} = 0
ery
due to the orthogonality condition for principal strains. In a completely
analogous fashion we have eys = ¢ = 0. This means that there is no
change in the right angles between the principal 2’, y’, and z’ axes as a
result of the deformation.
Consider an element of a strained medium which before deformation
occupies a volume V. Let the final deformed volume be Vp. The volu-104 Five
metric strain (or cubical dilatation or simply the dilatation) is defined as
_Vo-V
a- 2 (2.17)
In particular, let the volume element V be in the form of an infinitesimal
rectangular parallelepiped with its edges in the direction of the principal
axes of strain. Upon deformation, this element becomes again a rec-
tangular parallelepiped. Let ¢1, ex, es denote the strains of the elements
in the principal directions. Then
Vo = (1 + es) + e)(1 + es) 0
so that from (2.17) the dilatation is given by
A= e1 + €2 + €3 + e1€2 + €2€3 + e301 + e1€0€3 (2.18)
3 Theory of Small Displacements
The formulas developed in the first two sections apply to so-called
states of finitestrain. They (or special modifications of them) are required
to accommodate large displacements of thin plates, shells, and columns
for both nonbuckling and buckling problems. However, many problems
in the theory of elasticity and plasticity involve very small (or infini-
tesimal) deformations, so that squares or products of quantities are neg-
lected in, comparison with the first powers. ‘Thus, for example, the strain-
displacement relations (1.4) are replaced by the linear approximations
Gee = Ur Gy = Wy Gee = We
fyz = E(u, + 02)
40. + wy)
fez = Css = (Ws + Us)
(3.1)
Also, no distinction is made between the e’s and the e quantities.
Therefore, for example, Eq. (1.8) is replaced by e: = és. This means
that the term e,*/2 is neglected in comparison with e:.
Similarly, the dilatation A as given by Eq. (2.18) for finite strains
reduces to
Atatatasateta e+ 2447, G2)
Therefore the dilatation is simply the divergence of the displacement field.
Now, given a displacement field with differentiable (and therefore
continuous) components u(2,y,2), v(z,y,2), w(2,y,2) it is a simple matter to
calculate the corresponding strain components by substitution into Eq.
(8.1). If, however, we try to solve the converse problem of first selectingthe six components of strain in an arbitrary fashion, it soon becomes
apparent that in general an associated displacement field cannot be found.
It can be shown that for simply connected bodies a single-valued displace-
ment field which satisfies (3.1) exists if and only if the strain components
satisfy the following diJerential equations of compatibility.
Sey , O%er _ 2é%exy
oy? da dy
Or, Dey _ 26%eys
ay? | dz? ~ dy dz
dere y Pew _ 26%ere
et * Ox? ~ Oe ac (3.3)
Bere | Oye _ ere , ry "
dy dz" Ox? ~ dz dy * dz dx
Dey 1 Ieee _ 9% Weve
dzdx' dy? dydz dx dy
op Mery Pee 5 See
a2 dean" Oy a
It may be recalled that a region of space Ris said to be simply connected,
or, more briefly, simple, if and only if every closed curve lying in R can be
shrunk continuously to a point without passing outside the boundaries of
the region. Thus the region between two concentric spheres is simple,
while the interior of a torus (ie., a doughnut) is multiply connected.
Similarly, the interior of a rectangle is a planar simply connected region,
while the annular region between two concentric circles is multiply
connected (more specifically, doubly connected). It should be noted
that additional conditions to (3.3) must be satisfied for multiply connected
bodies.
Example 1: An elastic solid of arbitrary shape is subjected to a
temperature distribution of the form T = T(z,y,2), where aye is a
Cartesian system of coordinates. If each element in the body is free
to expand (i.e., the body is not stressed), the strain components are
fez = Gy = Gee = aT (3.4)
fy = G2 = Gz = 0
where « is the linear coefficient of thermal expansion and is assumed to be
constant. Determine the necessary form of the temperature distribution.
Solution: If (3.4) is substituted into (3.3), there résults
Tz + Ty =0
Ty + Tee
Ti t+ Tiz =0
Tye = Tez = Ty =0
(3.5)106 Five
which implies that
Pex = Ty = Te = Ty = Tye = Ter = 0
Since all the second partial derivatives with respect to the space variables
are zero, the temperature field is at most a linear function of z, y, and z,
that is, T = A + Be + Cy + Dz, where A, B, C, D are constants. It
may be shown conversely that if the temperature distribution in a simply
or multiply connected body is linear in rectangular z, y, 2 coordinates and
there are no surface forces, body forces, or surfaces of displacement dis-
continuity, then there will be no stresses throughout the body.
Example 2: A procedure for determining the displacement field is to
first determine the strain field and then integrate the strain-displacement
relations. The rigid-body displacements are usually disregarded in these
calculations. Therefore, it is of interest to determine the form of the
rigid-body displacements corresponding to the linear strain-displacement
relations (3.1). The rigidity requirements imply that all components of
strain are zero. From ez = ¢y = é = 0, it follows that u, v, w are
independent of =, y, z, respectively. Therefore
u = uoly,z) v = v0(Z,2) w = wolz,y)
If these functional forms are substituted into the expressions for the shear
strains, we have
AuUo , Wo _ Oo | IW _ wo
Spe a eae
du
oz
If the first equation is differentiated partially with respect to y and it is
noted that vo does not depend on y, then
a :
a =0 @
Similarly if the third equation is differentiated with respect to z, there
results
au, i
a 79 wi)
‘The most general solution of (i) and (ii) is the bilinear form
tug = Aye + ay + bz + Cy
Similarly,
v9 = Bex + dx +ez+Cr
wo = Cay + gx + hy + Cswhere all coefficients are arbitrary constants thus far. Substitution into
dup , a
Gy t Ge 7 0 yields
(Az +a) + Bze+d=0
or A +B=Oanda+d=0. Similarly,
B+C=0 and e+h=0
C+A=0 and g+b=0
so that
Therefore,
U = Up = we — wy + Cr
v = 9 = wt — oy + Cr
w= Wo = wry — wot + Cs
If we let
CH=CitGjt+Ck r=aityjtck
© = ai + wx + wk
and
S = ui + oj + uk
there results the vector relation
S=C+exr (3.6)
Thus the displacement vector S consists of a translation C of the body as a
whole and a rotation. Of course, this agrees with the fact that the most
general displacement of a rigid body is of this form.
4 Stress
The concept of stress arises from an analysis of the internal forces
which act in a body. Suppose that we have a three-dimensional solid
body subjected to the action of surface and body forces. Select. a point 0
in its interior and pass a plane through O (Fig. 5.2). Consider the forces
exerted by one part of the body B: on the remainder Bz with the plane as
part of its boundary. Surround the point O with a closed carve in the
plane of area AA, The forces acting on AA may be replaced by a resultant
force AP at O and « couple AC. Let n be a unit outward normal to AA.108 Five
PLANE SECTION
THROUGH POINT O
EXTERNAL
FORCES ON BODY:
Fig. 5.2 External and internal forces
acting on and within a body, respectively.
By definition, the average force per unit area or average traction at O
on an area with normal n is then AP/AA and the average moment is
AC/AA. Let AA tend to zero, so that O remains inside AA and all its
linear dimensions tend to zero. Assume that the limit exists independ-
ently of the manner in which AA approaches zero. The traction vector or
stress vector is defined by
_ AP
Tew jim ad oY
Furthermore, it is usually assumed that |AC| is of the same order of magni-
tude as |AP| multiplied by a representative moment arm whose length lies
within the shrinking area AA. Thus while AP/AA does not in general
become zero in the limit as AA approaches zero, it will be assumed that
Ac
aan BA 9 (2)
In effect, we are not accommodating point moments, which could occur
when an external magnetic field acts on a magnetized medium.
There are several facts concerning the stress vector T, that are impor-
tant to note, namely,
1. T,, in general, is not parallel to the normal n.
2. T, is not only a function of position but also depends upon the
orientation of the planar element (or equivalently, the normal n).
3. If T, refers to the action of By on Bz, then we denote T_, as theaction of B:on By. But, since action and reaction are equal and opposite
(Newton's third law),
T..=—-T, (4.3)
In particular, for the tractions T., T,, T; on elements perpendicular
to the z, y, and z directions, respectively, we have
(4.4)
Since the traction T,, is a vector, it may be resolved into two com-
ponents, one perpendicular to the planar clement and the other in the
plane. We may write
Tr = ont + ons (4.5)
where n and s are unit vectors perpendicular to and in the planar element,
respectively. The quantity cnn is called the normal or direct stress com-
ponent and is positive when tensile and negative if compressive (see Fig.
5.3). Also, ons is referred to as the component of shear, since it tends to
cause relative sliding of one element along a contiguous element.
Alternatively, T, may be resolved into its Cartesian components. If
i, j, k denote three mutually orthogonal unit vectors in the 2, y, 2 direc-
tions, respectively, then let
= Ons + Omyj + onk 46)
and, in particular, if m = i, j, and k in turn, there results
Gack + onj + oak
Oval + onj + ok (4.7)
Te = Gest + nj + ook
2 A normal stress component is tensile if it is directed away from the surface upon
which it acts. If it is directed toward the surface, it is compressive.
Sanh Ty
Fig. 5.3 Resolution of traction vec-
Tm8 tor into orthogonal normal and
shear components.
AA110 Five
5.4 Cartesian components of
traction vectors acting on a rec-
tangular parallelepiped showing the
positive directions.
The nine Cartesian scalar components gzz, gz, - . - , dz: are shown in
Fig. 5.4. There ozs, oy, and gz, are the normal or direct-stress com-
ponents and oxy, ozs, Gye) Ty; Jzz; Jxy are the shearing components. The
meaning of the subscripts zz, zy, ... , zz is that the first subscript
indicates the coordinate axis normal to the element of area upon which
the traction acts and the second subscript indicates the direction of the
component of the traction vector. Therefore, for example, o., represents
the shearing component in the z direction of the traction vector T, acting
on a planar element z = constant. Also, a tensile normal component
implies that it is positive and a compressive normal component implies
that it is negative.
5 Relation among Traction Vectors at a Point
In the preceding section it is shown that the traction vector varies not
only with position but also with orientation of the element of area AA.
We now seek a useful relation among the traction vectors at a point from
the condition of force equilibrium.
Let OEDG be an elemental tetrahedron and h the distance from 0 to
triangle DEG (Fig. 5.5). Also, let AA, AAs, Ay, AA, represents the
areas of the triangles DEG, OEG, ODG, and ODE, respectively. The
unit normal n from the origin to the plane triangular element is assumed
to be oriented with direction cosines J, m, n referred to the Cartesian
system so that
n= Li + mj + nk (5.1)and
AA, =1AA AAy=mAA AA, = AA (5.2)
There are in general two kinds of forces acting on the tetrahedron:
1. Surface forces acting on its boundary. The resultants are given by
AAP,, AA,P_., Ady P_,, 44, P_, on the respective faces DEG, OBG,
ODG, and ODE.
2. Body forces distributed over the volume of the body such as gravi-
tational forces. Let F denote the average body force per unit of volume
(i.e., average body force density).
The vector equation of force equilibrium is therefore
P_.AA, + P_, AA, + P_,AA, + P,AA + 3h AA F =0 (5.3)
Division by AA and application of (5.2) and (4.4) yields
IP. + mP, + nP, = P, — $hF (6.4)
If we let h— 0, then P, > T., P, > T,, P. > T:, Ps Ts, the tractions
at O (assuming these tractions are continuous functions of position). In
the limit there results
T, = IT. + mT, + nT, (5.5)
If we then substitute (5.5) into (4.6) and (4.7) and equate coefficients
of i, j, k,
One = laze + Moye + Nose
Ony = lasy + mo yy + Roy (5.6)
One = laze + Moye + Noe
DAP
Fig. 5.5 Surface forces acting on
Aa,P, elemental tetrahedron of body.uz Five
Finally, from (4.5), (5.5) and (4.7)
oon = n+ T, = n+ (IT, + mT, + nT.)
= UGzel + 24m + o2n)
+ m(gyal + oyym + oyn)
+ (rel + 0.4m + oun)
or
Onn = Poze + Moy + Woe + lm(Gxy + oy2)
+ mn(oys + Gey) + MUGre + oz2) (5.7)
Also, let the unit vector s in (4.5) be given by
s=li+mj+n'k (6.8)
From (4.5), (5.5) and (4.7) we have
s-T, = s- (IT. + mT, + nT.)
Is+ (Gzsi + onj + onk)
$ms+ (6 + n§ + ok)
$.ns+ (Casi + onj + onsk)
Finally, from (5.8) there results
Tne
one = Woe + mm'oyy + nN’ + Im'oe,
t+ Umoy, + mn!oy: + non (5.9)
+ alors + n'los
Equations (5.7) and (5.9) give the normal and shear components of
traction in terms of the nine Cartesian components of stress.
6 Equilibrium Equations
Consider a body which is in a deformed state of equilibrium due to the
action of both body and surface forces (Fig. 5.6). For equilibrium, it is
necessary that the vector resultant of all the forces and moments about an
arbitrary origin be zero. Let T, denote the traction at the surface S, F
the body force per unit of volume (assumed to be continuous over V),
s TqdS
Fig. 5.6 Body and surface forces
acting on deformed body.and r the position vector from the arbitrary origin O to an element of
volume dV and surface dS. Then for force and moment equilibrium
f,T»a8 + [Fav =0 6.1)
and
far xTadS + [rx Fav =0 (6.2)
We shall now show that (6.1) yields the stress equations of equilibrium and
(6.2) will lead to the equality of cross shears, that is on = oan
Substitution of (5.5) into (6.1) results in
f, (Ts + mT, +n, as + [Fav =0 (6.3)
and if the divergence theorem of the vector calculus is applied (assuming
that the traction T,, possesses continuous first derivatives), then
eT at, cat, _
(E+ +B +r)av-0 64)
Equation (6.4) must hold for all volumes, i.e., for any proper subvolume
of the body. Therefore, since the integrand is a continuous function of
a
ot: 4 Oy at + Ee
aT.
a
which is a vector equation of equilibrium. Finally, substitution of
(6.5) into (4.7) and equating coefficients of i, j, and k to zero gives the
corresponding scalar form
dors
+F=0 (6.5)
oye dee
ae + Ses Met = 0
Sout 4 Sn 4, Son ta +F,=0 Co)
i i: se : ie a
where
F = iF. +3f, + KP, 6.2)
Next, consider the consequence of the vanishing of the resultant
moment produced by the body and surface forces. From (6.2) and (5.5),
Jar x (Ts + mT, +n) aS + [pr x Fav =0 (68)
Application of Gauss’ pa theorem to the first integral implies that
he @xT) +5 OxT) +2 @xT) +rxFlav =0
(6.9)14 Five
If we carry out the differentiation of (6.9), there results
Jprx «(E49 eons one + 2 av
“i xT + ExT, +ExT,)av =o (6.10)
However, the integrand of the first vote is zero in accordance with
(6.5), and furthermore, since (6.10) must hold for arbitrary volume V,
the second a is zero, that is,
ext +S ene kenoe (6.11)
or or
But 5. =i, 5, hs 5s oe = k, and consequently (6.11) reduces to
ixTe+)xTtkxT.=0 (6.12)
Insertion of (6.12) into (4.7) and equating the coefficients of i, j, and k to
zero yields the equality of cross shears, that is
Oxy = Fyz Oye = eye = ee (6.13)
Therefore, the state of stress in a body is determined by the six com-
ponents of stress o2z, Cy) Gs2) Tay = Sys) Sys = Oxy Ore = Tze.
More generally, it is a simple matter to show that
Ons = Ten (6.14)
All that is required is the interchange of the roles of n and s in addition to
(m,n) and (U'ym’,n’) in (5.9). Then apply (5.9) and (6.13) to yield (6.14).
7 Principal Stress at a Point
In Secs. 5 and 6 the relationships between the stress components at a
point were obtained. In particular, the normal component of stress nn
as given by (5.7) was expressed in terms of the Cartesian components
and the direction cosines (/,m,n) of the normal with respect to the zyz
system of coordinates. It is seen that with oz) = oyz, oy = Ory, and
Gzz = xz the relation (5.7) is completely analogous to the corresponding
formulas for the transformation of strain such as given by the first of
Eqs. (2.3). Thus (omitting the unnecessary details) it follows that
there must be three mutually orthogonal directions for which the normal
components of stress are stationary. ‘These directions are known as the
principal directions, and the corresponding stresses are termed the prin-
cipal stresses. The normal stress along two of the axes will actually be
the largest and smallest normal stresses, while along the third axis wewill have neither a maximum or minimum. Let «, o2, os denote the
principal normal stresses at the point. The planes normal to the prin-
cipal directions are called the principal planes of stress. The principal
stresses o are determined from the determinantal equation [analogous
to (2.8)]
Gx Oxy on
oy Owe Oe | =0 (7.1)
Ox on on — 8
Expansion of the determinant results in the cubic equation
o — Jy? + Ja — Js = 0 (7.2)
where Ji, J2, and Js are stress invariants defined by
Ji = G22 + Oy + on
Se = Czy + Oyee + on22 — Tny'
Tze Try Tze
Js=|ou ow ow
Oss Tye Oe
2 2
= Ons" (7.3)
Oy
Furthermore, the three roots of (7.2) must be real. Finally, it is very
easy to establish that the shear components of stress are zero on the
principal planes, i.e., the traction vector on a principal plane is per-
pendicular to the plane. For from (5.5) and (4.7), Ts = om is equivalent
to the system of equations (a = li + mj + nk)
Uox2 — 0) + mow + Now = 0
losy + My — 0) + oye
loss + Moy + N(Ge2 — 0) = 0
(7.4)
This is the system of equations which, for nontrivial [m,n (recall
1? + m? + n? = 1), leads to the determinantal equation (7.1) in exactly
the same manner as (2.7) implied (2.8). Thus the proof is complete.
Note that the stress invariants are related to the principal stresses by the
relations
Jv=atortos
Jy = a102 + 0105 + om (7.5)
Js = 10203
8 Hooke’s Law
The relationships among the strain components in Secs. 1 to 3 and then
the stress components in Secs. 4 to 7 are applicable to liquids and gases
as well as solids provided that the bodies may be represented as a con-116 Five
tinuum. The elastic properties of materials have thus far not entered
into the relations deduced up to this point. However, the interrelation
between the components of stress and
strain is based primarily on experimen-
tal data varying from material to ma-
terial.
This and ensuing sections of this
chapter are concerned with Linear elastic
solid bodies. The general meaning of
the word “elastic” is that there is an
absence of any permanent deformation
when a specimen is loaded and then
unloaded. The term “linearly elastic”
implies that the relationship between
“load” and “deflection” is a linear one.
For example, the load may actually bea
Fig. 5.7 Shaft twisted by twist- twisting moment while the deflection is
ing moment or torque. the corresponding rotation. Thus if a
steel shaft is subjected to a twisting
couple as shown in Fig. 5.7, a typical response curve of moment: vs.
rotation will be as shown in Fig. 5.8. It is noted that the twisting
moment M is directly proportional to the angle of twist 0 until the pro-
portional limit is reached. This defines the linear elastic response range.
FRACTURE
TWISTING MOMENT M
INELASTIC NONLINEAR RESPONSE
PROPORTIONAL LIMIT
LINEAR ELASTIC RESPONSE
° ANGLE OF TWIST 6
Fig. 5.8 Representative moment—angular displacement
curve.The greatest moment that can be applied without producing a per-
manent angle of twist upon unloading is called the elastic limit. For steel
the elastic limit very nearly coincides with the proportional limit, and
the distinction between the two may be disregarded. The range of
moment beyond the elastic limit defines the inelastic nonlinear response.
Note that a point in the loading process is reached after which the loading
actually reduces in the neighborhood of material fracture.
Similarly, consider a slender steel bar of initial cross-sectional area A
subjested to an increasing tensile load P (Fig. 5.9). If the load is assumed
to be distributed uniformly over the cross section and if small changes
are neglected in the cross-sectional dimensions, then the longitudinal
stress is given by oz: = P/A, while the longitudinal strain is ee = AL/L
(the change in length per unit of length in the rod), Hooke in 1678
postulated that the axial stress 0. is proportional to the strain; that is
Ow = Bess (8.1)
where E is the constant of proportionality and is known as Young’s
modulus. Equation (8.1) is known as Hooke’s law. It is essentially
valid for many materials until the proportional limit is reached. (Some
materials exhibit only an approximately linear stress-strain curve.)
9 Generalized Hooke’s Law
In our discussion of a rod under a simple uniaxial tension, we neglected
the change in the cross-sectional dimensions. When a rubber band is
stretched, it is easily verified that the dimensions of the cross section are
decreasing. In other words, an axial stress causes strains not only in
the same direction but also in the transverse direction. A logical gener-
alization of Hooke’s law (due to Cauchy) is to assume that at each point
of the medium the stress components are linear functions of all the strain
components and conversely. The resulting law is called the generalized
Hooke’s law and is expressed mathematically (for an isothermal state) as
G22 = Caress + Crreyy + Crates + 2Crueys + 2Crseee + 2Creeay
Caress + Cortyy + Crates + 2C240y2 + 2Cas€es + 2Coceey (9.1)
Fig. 5.9 Slender rod under axial tension.118 Five
The 36 coefficients Cu, Cis, . . . , Cesin (9.1) in general may be functions
of position. If, however, they are constants which are independent of
position, the medium is defined to be elastically homogeneous. We shall
from this point restrict our considerations to elastically homogeneous
systems and denote the 36 constants Cu, Cuz, . . . » Ces as the elastic
constants of the material. Also, a body is said to be isotropic if its elastic
properties are the same in all directions. What this amounts to is that,
for an isotropic body, the stress-strain relations referred to any orthogonal
Oz'y'z' coordinates must have the same coefficients as in (9.1). In other
words, for arbitrary orientation of the z’y'z’ Cartesian coordinate system
with respect to the zyz system the following will be the stress-strain law:
ate! = Creare + Cireyry + Crseerer
H+ Waser + Wrsesrer + WCreesry
yy = Crrezrar + Cortyys + Crsterer
+ WCreeyer + Woaseere + Woreesry (9.2)
ary = Corearr + Cortyy + Costes
H+ 2Coseyrer + Wester + Woeeery
where, it must be emphasized, the coefficients in (9.2) must be identical
with the coefficients in (9.1). It may be proved that for an isotropic
body the 36 constants are not independent of each other. In fact, the
number of elastic constants reduces to only 2. This may be established
by using the invariance of the stress-strain relations with respect to
various covering operations such as (1) a rotation about a line or axis and
(2) a reflection in a plane. ‘The detailed derivation of the reduction from
the 36 constants to 2 constants will not be given in this book. Instead,
the following facts will be noted:
1. It can be shown that with the assumption of the existence of a strain
energy function that Cy = Cx, ij, and i,j=1,2,...,6. This
reduces the number of elastic constants to 21 in the generalized Hooke’s
law and applies to the most general case of an anisotropic elastie body.
Crystals possessing no symmetry at all define the triclinic system.
2. If the medium is symmetric with respect to a plane, then the con-
dition of the invariance of C;; results in a reduction to 13 constants. ‘This
is called the monoclinic class.
3. For a material with three mutually orthogonal planes of elastic
symmetry, the number of elastic constants is reduced to 9. Such
1 For detailed discussions the reader is referred to (1) I. S. Sokolnikoff, “Mathe-
matical Theory of Elasticity,” 2d ed., pp. 56-70, McGraw-Hill Book Company, New
York, 1956 and (2) A. E. H. Love, “A Treatise on the Mathematical Theory of
Elasticity,” 4th ed., chap. 6, Dover Publications, Inc., New York, 1944.materials as wood possess these properties and are said to be orthotropic.
4. Further symmetry requirements of independence of the stress-strain
laws with respect to one pair of the mutually orthogonal reference axes
result in the hexagonal erystal class with 6 elastic constants (for example,
quartz). If, in addition, the same holds true for a second pair of axes, we
have the so-called face-centered or body-centered cubic class exemplified
by crystals of the majority of pure metals—iron, copper, nickel, and
aluminum, for example. It is readily shown that the number of elastic
constants reduces to 3 for this class.
While it is true that for a crystalline body the elastic properties depend
on orientation of axes, metals on a macroscopic scale may be considered
as isotropic. That is because many metals have a sufficiently small and
random-oriented crystal grain structure. For an isotropic body the
number of elastic constants reduces to 2 and the stress-strain law may be
expressed as
Gx2 =O + Qwese — oy = OT Bey oe = NOT Byles
Oye = Queye — Ore = Were ay = Aplery
(9.3)
where 8 = exe + éy + er is called the volumetric strain or dilatation and
the constants \ and u are called Lamé constants.
Conversely, (9.3) may be solved for the strain components in terms of
the stress components:
— Ot Wore _ x
“WGN 2a) — Bal -F Bay Cow + en)
CeO ee
HBX + 2y) ~ Bu(BN -F Bp) Me 2)
= Ot Won 2
»
“* ~ WGNF Bu) ~ TaGr Fo—) Cs + em)
ou ne
wi ee ae
10 Significance of Elastic Constants for Isotropic Materials
Very important physical insight may be gained by considering the
response of elastic isotropic materials to particular loadings which result
in states of pure shear, uniaxial tension, and hydrostatic compression,
respectively.
Pure Shear. We may characterize this state by the requirements
x = % = T = constant throughout the body while all the rest of the
stress components are zero. From (9.4) it follows that correspondingly
there is only one nonzero strain component éy = é2 = T/2u. Thus in120 Five
accordance with Sec. 1, we may interpret » = 7'/2ez, as the shear stress 7’
divided by the decrease in angle between linear elements originally
parallel to the z and y axes (Fig. 5.10). The quantity wis called the shear
‘modulus and in the usual stress analysis is denoted by the letter G, that is,
aS eB (10.1)
and is a constant which assumes different values for different materials.
Uniaxial Stress Field. Next let us return to the slender cylindrical
bar of Fig. 5.9 subjected to the action of longitudinal forces applied at the
ends. If the load is taken to be uniformly distributed over the cross
section, then the stress field will be o,, = constant while the other com-
ponents are zero. Then from (9.4), the direct strains are
Obes
= (Bd + 2p)
= Mas
fa = © Dux + 2u)
(10.2)
But for the uniaxial stress distribution, we define Young’s modulus EB by
the formula E = ¢z:/éz. Thus from (10.2),
y= He = 2)
moe (10.3)
This relates i, #, and Z, but \ has no physical significance. Elimination
of ozz from Eqs. (10.2) results in
Deze
aL) = Te
oo SS 2H)
Fig. 5.10 Pure shear—rectangular
element ABCD before deformation
becomes a parallelogram AB'C'D
T after deformation with the samewhere » is defined by
x
Secret 10.4
FH) oe
‘The constant », called Poisson’s ratio, is the negative of the ratio of lateral
strain to longitudinal strain when subjected to uniaxial longitudinal stress.
Hydrostatic Compression. In the case of » uniform hydrostatic
pressure of amount P, we have
=-P
Oy = Os = oy = 0
Osx = Oy =
Then from the stress-strain relations
eee
Gee ioe oe eae ce Skat
or solving for P,
P= ere |
where K is defined by
Ke= (10.5)
The constant K, called the bulk modulus of elasticity (or modulus of
volume expansion), is the ratio of applied hydrostatic pressure to the
change in volume per unit of volume. By means of (10.1), (10.3), (10.4),
and (10.5) it is seen that any one of the elastic constants E, », @, K, d, «
may be expressed in terms of any other two constants, i.e., only two
constants are independent.
If quantities E and y are selected as the two basic elastic constants,
Eqs. (9.4) and (9.3) assume the more usual form
es = Fae — 1 + 60))
en = Flew — 1a + on]
fe = flow — (ou + on)]
ys ui (10.6)
l+p
fee = ee122
Five
or, conversely,
E
ee = ET Ml — eee + olen +
—. wise [0 = rey + r( eee + €ee)]
E
2 = BID Ih = vee + v(ces + ew)]
ey v)(1 — 2v) (10.7)
ee ee eri
E
oe re
ace
oe = TG
From their physical interpretations, the elastic constants u or @, B, and K
must be positive quantities. But from w= 3K(1 — 2»)|2(1 +») it
follows that the values of Poisson’s ratio are bracketed by —1 < » <}.
The following table gives average values of H and u obtained experi-
mentally for some materials in this linear clastic range. Note that the
Walues of ¢ arc deduced from y= z —1, although they may also be
determined experimentally. ‘The values of Z and u are given in millions
of pounds per square inch, and y is nondimensional.
Material
Aluminum alloy
0.313
Carbon steels 0.283
Cast iron 0.269
Glass 0.250
Magnesium 0.354
Malleable iron 0.269
‘Titanium
A very important consequence of the stress-strain law (10.6) or its equiv-
alent (10.7) is that, for isotropic elastic materials, the principal azes of
stress and strain coincide. This follows from the simple argument that if
the 2, y, and z axes are chosen along the principal axes of stress, then
Ory = Oye = G2 = 0. But from (10.6), this implies that ex, = ¢y2 = é2 = 0.
Therefore, the z, y, and z axes are also principal axes of strain. This
coincidence of the principal axes of stress and strain for isotropic bodies
may be employed to derive (in reverse) the stress-strain relations using theconcept of superposition. If we imagine a rectangular parallelepiped
element of the body subject to the simultaneous action of normal stresses
in the three coordinate directions, then from the definition of Z and »,
= G->(B+9)
Similar results hold for ¢,, and és, showing agreement with (10.6). This
approach is not as satisfying from a theoretical viewpoint. However, it
does appeal strongly to our physical intuition.
11 Strain Energy
Consider a body acted upon by surface tractions T, and body forces F
per unit volume. If the change in the body heat is considered negligible,
and assuming that the body is in a statically deformed state, then the
work done by the external forces must equal the strain energy stored in
the material.
Let U denote the strain energy of the material and Us denote the strain
energy per unit of volume, so that
u=f, Ueav (uy)
where the integration is performed over the volume of the body.
Suppose that the displacement field u undergoes a change from
u to u + du (a virtual displacement compatible with the constraints).
The corresponding work done by the surface tractions is [, T,+ du dS,
where the integration is taken over the surface of the body. Similarly,
the work done by the body force F per unit of volume is i: F- éudv.
Thus, the balance of energy requires that
J, Ueav = [,T.- duds +f, F-ouav
= [, (Ts + mT, + nT) suds + f,P-suaVv
(11.2)
where (l,m,n) are the direction cosines of the unit outward normal n to
the surface. Application of the divergence theorem to the first integral
on the right side of (11.2) yields
a a
J, avvav = f,[Z ow + Z cry ou)
a
+56
av + [,F-suav124 Five
which upon expanding becomes
aT , aT, , aT
J, eUeav = i (E+ Fe SE yw). amar
yu du du
+f, [te 040 oe Ela (11.3)
But
oT.
Ox
oT, me
oy
+F=0
from (6.5), so that
|, sUeav = ffm +138 ET, 62 ]av (14)
From (4.7) and the expressions (3.1) for the strains in terms of displace-
ments we have
fi, 8U0aV = f, loxsbeee + Obey + ober
+ Qondery + 2oysbeys + e262] dV (11.5)
If it is assumed that the strain energy density is a function of the strain
components, that is,
Uo = Uol€szs€vvy6s2s€zys€ves€zz)
then
Uo 4. 4 Ue eee aus
jf, oUeav = i [se Bee + Ge Sew + + Fee] av
(11.6)
The right sides of (11.5) and (11.6) may be equated. This equality must
hold for arbitrary volumes and, furthermore, des, dé, . . - , dex are
independent variables to within equilibrium and compatibility require-
ments. Therefore, we deduce the equalities
, = We _aUy _ ao
at es aasae ess ges
aU _ Uo aUo a.
Cele open ee Ogee odes
Finally, the significant expression for the strain energy density Uo can
be obtained by expressing it as a homogeneous quadratic function of the
components of strain (as a trial function). Then, utilization of (11.7)and the linear stress-strain law for isotropic bodies (9.3) yields
Uo = ATH esata? He) + Menten + tant + ete)
A Qu (exy? + eve? + €rs*) (11.8)
which may be rewritten in the form
= xe Hh Mlean* + yy? + eee” + lay? + 2eys® + ees”) (11.9)
Since 4 > 0, that is, » > 0 for any physical material, it follows that the
strain energy density Up is a positive definite form in the strain components.
Substituting Eqs. (10.6) into (11.9) and using the expressions for \ and yz
in terms of H and », the function Uy is expressible in terms of the stress
components as
Uo = iy least + om? + 0n8 — 29(Cyytee + orsbee + ons0)
$20 + Gon? + on2 + on)] (11.10)
and this is also a positive definite form in the stress components, since
vcd
Also, Uo may be expressed in terms of both stresses and strains as
Uo = Bosstes + ytiy + Gestes + 2ozyery + yeeys + rates] (11.11)
Finally, if the expression (11.10) is differentiated partially with respect
to each of the stress components and use is made of the linear stress-
strain relations (10.6), we have
sth ay | _ ats
"Gee lon | A ae
aU> aU. Us ae
elgg ee oaze | oe a ageSix
Energy Principles,
Methods, and Applications
1 Theorem of Minimum Potential Energy
In the preceding chapter we employed the principle [expressed by (11.2)]
that the virtual work performed by the body and surface forces during an
arbitrary variation in displacements is equal to the increment of strain
energy. Since the volume V is fixed and the surface tractions T, and
body forces F do not vary with arbitrary virtual displacements (com-
patible with constraints of system), (11.2) of Chap. 5 may be rewritten
in the form
a[f, vear - [,T.-uas - f,F-uav] =0 (a.
‘The expression in brackets is defined to be the potential energy V of
an elastic continuum, that is
v= f, Usa — [,T.-uas — f,F-ua? (1.2)
and (1.1) becomes simply
sv =0 (1.3)This states that the potential energy of an elastic system in equilibrium
must be stationary with respect to all displacements satisfying the given
boundary conditions. In fact, more strongly, not only must the potential
energy be stationary but it can be shown that it actually assumes a
minimum value when the displacement field is that of the stable equilib-
rium state. Conversely, it can be established that if the potential energy
of an elastic system in equilibrium is an absolute minimum for a displace-
ment field u which also satisfies the boundary conditions, then the equa-
tions of equilibrium must also be satisfied by the field. It is this converse
form that is most prevalent in applications. To illustrate the application
of the theorem we shall consider two examples.
Example 1: Deflections of an Elastic String. The principle of
minimum potential energy will now be applied to determine the ordinary
differential equation for the small deflections of a flexible string. A
string (Fig. 6.1) is shown loaded by a distributed vertical force of intensity
q(x). The string is of length L and is under a large tensile load S$ when
q(2) is applied. We neglect the small change in the tension due to the
distributed force q(z). It is further assumed that the string is perfectly
flexible, so that the tensile force is directed along the tangent.
‘The strain energy in the string is composed of two parts: (1) U’ due to
the strain from the original tensile force and (2) U’" due to the deflection
of the string resulting from the transverse loading. The total strain
energy is given by U = U’ + U”. Let us find an expression for U”.
‘The increase in strain energy due to the deflection is obtained by multi-
plying S by the increase in length due to the deflection. If y denotes the
string deflection,
ur = 8 fo (as— ae) = 8 [2 [ahi + (#) - Jae
Small deflections are taken to mean that (dy/dz)*<< 1. Then according
Fig. 6.1 Vertically loaded string.
(a) String under tension S with no
(o) vertical load. (b) Same string with
vertical load added.128 Six
to the binomial expansion,
ae a rr S pz (ay i
U! = 8 f7[1+3 (2) + ee dz (i)
and the total strain energy is given by
_ oS pe (au? i
u=uU+5f (ya (i)
Substitution of (ii) into (1.1) for the ease of no body forces results in
cee ae L ue ae
ale +5), (2) ax — [ aye] = 0 (iii)
where — [(" ay dx is the potential energy of the distributed load q(a).
But U’ is constant for all variations in displacement, so that 5U’ = 0.
Thus (iii) reduces to
[5 (Be) a - faves] -0 (v)
Therefore the Euler equation corresponding to the integrand
S (ay)?
3 (Hy - 9
FY + g(a) =0 (4)
is
which is the differential equation of the vertically loaded string. Its
solution may be obtained by repeated integration where constants of
integration are determined from the boundary restraints (0) = y(L) = 0.
For example, in the case of a uniform load q = go = constant,
y= gee — 2) (1.8)
which is the well-known parabolic deflection shape.
Example 2: Beam Column. The beam column represents a general-
ization of the ordinary beam problem in that it is a beam subjected to
both axial and transverse forces. We shall be concerned with the deriva-
tion of the differential equation of equilibrium and boundary conditions
using variational notation (Chap. 3, Sec. 5) in conjunction with the
principle of minimum potential energy. Also, a solution will be developed
for the deflections from which the bending moment and stress may be
readily calculated.
Figure 6.2 shows the problem to be analyzed. The applied loadings
are compressive axial end forces P, distributed transverse loading q(z),and two concentrated end moments My and M,—all taken to be acting
in a principal plane of bending. Forces and moments are taken to be
positive as displayed, and the beam ends are simply supported for
bending.
‘The strain energy of bending is
U=sf Bry’) ae @
while the potential energy of the distributed load q(2) is
Vex — fl wae Gi)
Similarly, the potential energy of the moment Mo is
Vu, = —Moly')2-0 Gi)
while the corresponding quantity at the other end is
Vu, = Mily')enn (iv)
(Note that the rotation in the direction of Mz atz = Lis (—y’)ea1-)
For the compressive end loadings, the potential energy is — P times the
axial shortening. The latter quantity is found from the condition that,
during bending, the axial beam length does not change. This is a char-
acteristic of a ‘small-deflection” analysis. If A denotes the displacement
of the end x = L in the direction of the load P, then
[PO vIF Uae =
is required. But V1 + (y’)? = 1 + 3(y’), so that
fru +40 a = 7
or
Lats [wars
Fig. 6.2 Loads applied to beam column.130 Six
Finally, since A is negligible in comparison with L,
asa fl? wre sd [wae w)
Therefore,
Ve= 2 [i wae (vi)
The principle of stationary potential energy requires that
6(U + Va + Mo+ Mi + Vr) =0 (vii)
or from (i) to (iv) and (vi)
oe :
{ft [ere — Fw — a] de + Mada
— Maly} = 0 (vii)
If the variation is performed, we have
1
fh [EIy" by” — Py! by’ — q by] dx
+ M1 (3y')2nr, — Mo(5y')zno = 0 (ix)
Application of integration by parts twice yields
Jo Blu" by" de = (Bly ay’ — f= Ely") by de
4 1
= tery’ ov — [2 ty") ov |
ae
+ [C5 Ev) yar
a d miyt
= Wly" iy! + ff F5 ly") by de @)
while
i : d
Jo Py oy ax = Pry’ by - [2 ey) ay ae
aa, :
= — [PE Pv) iy ae xi)
since 6y = Oateach end. Substitution of (x) and (xi) into (ix) results in
LT da? = d -
kl fen + Zen ~ a] ee
+ (Mz + Ely'!) by! — (Mo + Ely}’) by’ = 0 (xii)Therefore, the required differential equation is
Oe
qa Ely") + (Py) —9=0 as)
and the boundary conditions are
(Mi + Ely) =0
(My + BIW = 0 eo
and
yruy=0 (1.8)
If q = Mo = Mz = 0 (that is, if there are no transverse loading or end
moments), then we have the eigenvalue problem for pin-ended columns
2
# ly") + Py’ =0 (1.9)
subject to the boundary restraints
Yo =O (y)zan = 0 (1.10)
and
y0)=0 y(L) =0 (ay)
If, in particular, the bending rigidity ET is a constant, then the problem is
further reduced to the Euler buckling problem.
Also, if P = 0 (no axial end loads) the differential equation (1.6)
becomes the beam equation,
£ Gy) -¢=0 (1.12)
Solution of the Boundary-value Problem. We now return to the
problem of solving the boundary-value problem (1.6) to (1.8) for the case
when the bending rigidity and the applied transverse loading q are
constant. Equation (1.6) then becomes
y+ hey! = (1.13)
where
P
woe (1.14)
Let y/’ = u, so that (1.13) becomes
ul + ku = H (1.18)
‘The solution of (1.15) is composed of any particular solution added to the132 Six
complete solution of the homogeneous equation w” + ku = 0. Therefore,
rae ae q
u = A’ cos ke + BY sin kx + Ses (1.16)
where A’ and B’ are arbitrary constants. Note that since cos ke and
sin kx are linearly independent functions," the solution is complete.
Two successive integrations of u yield
= i ge?
y = A cos ke + Bein ke + sia; + Ce +D (1.17)
where A = —A'/k? and B = —B’/k? are new arbitrary constants, while
C and D are constants of integration. These quantities are now to be
determined from the boundary conditions.
The deflection is zero at x = 0,
(yo =0=A+D (a)
Also [y’"Je0 = — Me, But y” = —KA cos kx — WB sin ke + pha
therefore
= ge Me
fh eee aeRT,
or
A-ae (é + Me) (b)
From (a) and (b) substituted into (1.17),
’- TH (™ + 4) (cos ke — 1) + Bsinkr + 52 4 Ce
(1.18)
and
y= ~ By (Mo + fh) eos he ~ Bit sin be + phy
BI B PBI
‘The satisfaction of the conditions [yJeur = 0 and [y"lez = — M¥ yiela
B and C, so that after some straightforward simplification
y= per (Mery B) (eos ke — 0
io ae
|
- [a -M+% by, - »]z (1.19)
The stresses may now be calculated in a routine manner.
‘The formal definition of linear independence of n functions is given in Chap. 7,
See. 1.Final Comments
1. For completeness, it may be remarked that beam columns supported
differently from the foregoing development can be accommodated with
equal facility. If, for example, the member is built in, the boundary
condition is on the slopes rather than the moments.
2. The above development is not valid if the transverse loads are con-
centrated at points along the beam length. This is due to the finite jumps
in the shearing force and therefore the third derivatives at such points.
The differential equation (1.6) is correct in between the loads, and this
problem can be treated by integrating in steps.
Example 3: Three-dimensional Displacement Equations of Elas-
ticity. An isotropic body obeys Hooke’s law relating stress and strain
components. It is in equilibrium under the action of body forees per unit
of volume (F.,F,,F.) and surface tractions (oa2,¢n,0x2). It is required to
find the small displacement differential equations of equilibrium for
(w,»,w) by application of the principle of stationary potential energy.
Before we proceed to the solution it should be remarked that the
following technique is utilized frequently for derivation of equations
in structural mechanics. The potential energy V from its definition and
(11.8) of Chap. 5 is given by
Ve I if [G + ) (eae? + ean? + east) + Nest + frites + Gentes)
+ 2ulea? +e + ee) — Pau + Fw + Fw] de dy dz
— ff nat + ony0 + onetd) dS (1.20)
is
Next, let us vary only wu, that is, 6» = 6 = 0. Thus
Us bere = Suz = (5U)z
vy bey = ov, =0
fe = Ws bes = bw, =
ty +02 buy _ (5u)
oe ea, = She = Oude
Us + Wy
Ge bee
2
wee BEM gg, = ue
The surface tractions enter only into the boundary conditions and do not
affect the equations of equilibrium. Therefore, the surface integral may
be disregarded for our purposes. Setting the variation of V equal to zero134 Siz
results in
If f {G + +) 2us( bux + M(vy + 2)(Bu)s}
+ wl(ty + v2) (5u)y + (te ++ we) (3u),] — Pe au de dy az
+ boundary terms = 0
Integration of terms involving (éu)., (6u)y, and (5u), by parts and
changing signs yields
Jf] (0 + 2eduae + Nye + tee) + lt + Om)
7
+ pluses + w22] + Fe} du dx dy dz + boundary conditions = 0
Since u is arbitrary, the differential equation is found by setting the
expression in brackets equal to zero. Rearrangement of terms yields
aVut +n) 2+ 7 =0
where
: 2 du, ov , dw
A = dilatation = ie + ay + 7
and
VU = ez + Uyy + Use
‘The other two equations of equilibrium may be determined by cyclical
permutation of the variables (u,v,w) and (z,y,z). Thus we have, by
eliminating \ from (10.4) of Chap. 5, the coupled system of equations
Oh oye
(1 = 29) Vu + 5 + a F,=0
oA , 1-2:
(= 29) ve + FF + FF, =0 (1.21)
oA, 1-2
Ge 2?) Vea iz F,=0
2 Hamilton’s Principle for an Elastic Continuum
In Chap. 4, Sec. 3, Hamilton’s principle for a discrete system was
derived for a conservative force field. It is natural to anticipate that this
concept may be extended to a continuum. Our objective is to establish
this result and then to show some examples illustrating its use in problems
of motion.