Sie sind auf Seite 1von 131

Introduction to Continuum Mechanics

c Romesh C. Batra, 1998, 2000

CONTENTS
Chapter 1 Introduction 1-1 1-2 1-3 1. What is Mechanics? 2. Continuum Mechanics 3. An example of an ad-hoc approach Chapter 2 Mathematical Preliminaries 2-1 2-3 2-4 2-6 2-7 2-9 2-12 2-19 Kinematics 3-1 3-3 3-5 3-6 3-9 3-11 3-14 3-21 3-24 3-33 3-35 3-38 3-45 3-50
ii

1. Summation convention, Dummy Indices 2. Free Indices 3. Kronecker Delta 4. Index Notation 5. Permutation Symbol 6. Manipulation with the Indicial Notation 7. Translation and Rotation of Coordinate Axes 8. Tensors Chapter 3

1. Description of Motion of a Continuum 2. Referential and Spatial Descriptions 3. Displacement Vector 4. Restrictions on Continuous Deformation of a Deformable Body 5. Material Derivative 6. Finding Acceleration of a Particle from a given Velocity Field 7. Deformation Gradient 8. Strain Tensors 9. Principal Strains 10. Deformation of Areas and Volumes 11. Mass Density. Equation of Continuity 12. Rate of Deformation 13. Polar Decomposition 14. Innitesimal Deformations

Chapter 4

The Stress Tensor 4-1 4-10 4-13 4-15 4-20

1. Kinetics of a Continuous Media 2. Boundary Conditions for the Stress Tensor 3. Nominal Stress Tensor 4. Transformation of Stress Tensor under Rotation of Axes 5. Principal Stresses. Maximum Shear Stress Chaper 5 The Linear Elastic Material

1. Introduction 2. Linear Elastic Solid. Hookean Material 3. Equations of the Innitesimal Theory of Elasticity 4. Principle of Superposition 5. A Uniqueness Theorem 6. Compatibility Equations Expressed in terms of the Stress Components for an Isotropic, Homogeneous, Linear, Elastic Solid 7. Some Examples. a) Vibration of an Innite Plate b) Torsion of a Circular Shaft c) Torsion of Non-Circular Cylinders Chaper 6 The Linear Elastic Material

5-1 5-1 5-7 5-10 5-1ll 51-4 5-18 5-22 5-26

1. Constitutive Relation 2. Formulation of an Initial-Boundary-Varlue Problem 3. Examples

6-1 6-3 6-4

iii

1 Introduction
The major objective of our study of Mechanics is the formulation and solution of initial-boundaryvalue problems that model as realistically as possible a physical phenomenon. There are two equally attractive approaches to Mechanics. One is the ad-hoc approach, which takes up specic problems, and devises problem-dependent methods of solution, introducing simplifying assumptions as needed. (This approach is used in Strength of Materials where problems of bending, torsion, pressure vessel are individually set up under varying assumptions and then solved.) The other is the general approach, which explores the general features of a concept or a theory and considers specic applications at a later stage. By and large, the latter is the quicker way to learn about an entire eld, but the former is more concrete and sometimes more easily understood. We will study the general approach in this course. 1.1 What is Mechanics? Mechanics is the study of the motion of matter and the forces required to cause its motion. Mechanics is based on the concepts of time, space, force, energy, and matter. A knowledge of mechanics is needed for the study of all branches of physics, chemistry, biology and engineering. The consideration of all aspects of mechanics would be too large a task for us. Instead, in this course, we shall study only the classical mechanics of non-polar continua. (A nonpolar continuum is one whose material particles have only three translational degrees of freedom.) We shall concern ourselves with the basic principles common to uids and solids. 1.2 Continuum Mechanics Matter is formed of molecules which in turn consist of atoms and subatomic particles. Thus matter is not continuous. However, there are many aspects of everyday experience regarding the behavior of materials, such as the amount of lengthening of a steel bar under the action of given forces, the rate of discharge of water in a pipe under a given pressure difference or the drag force experienced by a body moving in air etc., which can be described and predicted with theories that

pay no attention to the molecular structure of materials. The theory which describes relationships between gross phenomena, neglecting the structure of materials on a smaller scale, is known as the continuum theory. The continuum theory regards matter as indenitely divisible. Thus, within the theory, one accepts the idea of an innitesimal volume of material referred to as a particle in the continuum, and in every neighborhood of a particle there are always innitely many particles present. Whether the continuum theory is justied or not depends upon the given situation. For example, the molecular dimension of water is about 1 A

(10;8 cm); hence, if we are concerned

about the liquid water in a problem in which we never have to consider dimensions less than say

10;5 cm, we are safe to treat water as a continuum. The mean free path of the molecules of air on the surface of the earth at room temperature is about 5 10;6 cm; hence, if we consider the ow
of air about an airplane, we may treat air as a continuum. The diameter of a red blood cell in our body is about 8:5

10;4 cm; hence, we can treat our blood as a continuum if we consider the ow

in arteries of diameter say 0.5 mm. Thus the concept of a material continuum as a mathematical idealization of the real world is applicable to problems in which the ne structure of the matter can be ignored. When the consideration of ne structure is important, we should use principles of particle physics, statistical mechanics, or a theory of micropolar continuum. 1.3 An example of an ad-hoc approach. Consider the problem of the bending of a beam usually studied in the rst course on Mechanics of Deforms. This is generally based on the following assumptions: i) The beam is initially straight. ii) The cross-section is uniform. iii) The beam is made of a homogeneous and isotropic material which obeys Hookes law. iv) Plane sections remain plane. v) The beam is subjected to a pure bending moment M applied at the ends.
2

Under these assumptions, one can derive the formula

= My I
in which is the longitudinal bending stress,

(1.3.1)

y the distance from the neutral axis which passes

through the centroid of the cross-section and I the moment of inertia of the cross-section about the neutral axis. The derivation of (1.3.1) makes no reference to other components of stress acting at a point. Of course, if the beam were initially curved or were one interested in nding the transverse shear stress at a point, one would start essentially from scratch. 1.4 Continuum Mechanics In Continuum Mechanics, we rst establish principles which are applicable to all media, both uids and solids, under all kinds of loading conditions. We then study constitutive equations which dene classes of idealized materials. Finally specic problems are analyzed, and results are compared with experimental observations.

2 Mathematical Preliminaries
2.1 Summation Convention, Dummy Indices Consider the sum

s = a1 x1 + a2 x2 + : : : + anxn :
We can write it in a compact form as

(2.1.1)

s=
It is obvious that the index i

n X i=1

aixi =

n X j =1

aj xj =

n X m=1

am xm :

(2.1.2)

j or m in eqn. (2.1.2) is dummy in the sense that the sum is indepen-

dent of the letter used. This is analogous to the dummy variable in an integral of a function over a nite interval.
Z b Z b Z b

I=

f (x)dx =

f (y)dy =

f (t)dt :

We can further simplify the writing of eqn. (2.1.2) by adopting the following convention, sometimes known as Einsteins summation convention. Whenever an index is repeated once in the same term, it implies summation over the specied range of the index. Using the summation convention, eqn. (2.1.2) shortens to s = ai xi

= aj xj = am xm .

Note that expressions such as ai bi xi are not dened within this convention. That is, an index should never be repeated more than once in the same term for the summation convention to be used. Therefore, an expression of the form

n X i=1
must retain its summation sign.

aibi xi

In the following, unless otherwise specied, we shall always take n to be 3 so that, for example,

aixi = am xm = a1 x1 + a2x2 + a3 x3 aii = amm = a11 + a22 + a33 :


The summation convention can obviously be used to express a double sum, a triple sum, etc. For example, we can write

3 X 3 X

i=1 j =1

aij xixj

simply as aij xi xj . This expression equals the sum of nine terms:

aij xi xj = ai1 xix1 + ai2 xi x2 + ai3xi x3


= a11 x1 x1 + a21 x2 x1 + a31 x3 x1 + a12 x1 x2 + a22 x2 x2 + a32 x3 x2 + a13 x1 x3 + a23 x2 x3 + a33 x3 x3 :
Similarly, the triple sum

3 X 3 X 3 X

i=1 j =1 k=1 represents the sum of 27 terms.

aibj ck xi xj xk will simply be written as aibj ck xi xj xk , and it

We emphasize again that expressions such as aii xi xj xj or aij xi xj xi xj are not dened in the summation convention.

Exercise: Given

1 0 2 aij ] = 4 0 1 2 5 : 3 0 3 Evaluate (a) aii , (b) aij aij , and (c) ajk akj .
2.2 Free Indices Consider the following system of three equations:

y1 = a11 x1 + a12 x2 + a13 x3 = a1i xi y2 = a21 x1 + a22 x2 + a23 x3 = a2i xi y3 = a31 x1 + a32 x2 + a33 x3 = a3i xi:
These can be shortened to (2.2.1)

yi = aij xj i = 1 2 3:

(2.2.2)

An index which appears only once in each term of an equation such as the index i in eqn. (2.2.2) is called a free index. A free index takes on the integral number 1,2, or 3 one at a time. Thus eqn. (2.2.2) is a short way of writing three equations each having the sum of three terms on its right-hand side. Note that the free index appearing in every term of an equation must be the same. Thus

ai = bj
is a meaningless equation. However, the following equations are meaningful.

ai + ki = ci ai + bicj dj = 0:
If there are two free indices appearing in an equation such as

Tij = AimAjm i = 1 2 3 j = 1 2 3

(2.2.3)

then it is a short way of writing 9 equations. For example, eqn. (2.2.3) represents 9 equations; each

one has the sum of 3 terms on the right-hand side. In fact

T11 = A1m A1m = A11 A11 + A12 A12 + A13A13 T12 = A1m A2m = A11 A21 + A12 A22 + A13A23 T13 = A1m A3m = A11 A31 + A12 A32 + A13A33

........................................... ........................................... T33 = A3m A3m = A31 A31 + A32 A32 + A33A33 : Again, equations such as

Tij = Tjk Ti` = Aim A``


are meaningless. 2.3 Kronecker Delta. The Kronecker Delta, denoted by ij , is dened as

ij
That is,

=
33 21

1 if i = j 0 if i 6= j: =1 =
23 12 22 32

11 12
In other words, the matrix

= =

22 13

= =
2

31

32

= 0:

11 4 21 31
is the identity matrix
2 4

13 23 5 33
3

1 0 0 0 1 0 5: 0 0 1 =1+1+1=3 = a1 = a2 = a3 :

We note the following relations (a) ii

11 + 22 + 33

(b) 1m am

= = =

11 a1 + 12 a2 + 13 a3 21 a1 + 22 a2 + 23 a3 31 a1 + 32 a2 + 33 a3
6

2m am 3m am

Or, in general

im am
Similarly, one can show that

= ai : = Tij :

im Tmj
In particular,

im mj
2.4 Index Notation

ij

im mj jn = in :

Usually, rectangular Cartesian coordinates of a point are denoted by (x along x

y z) and the unit vectors

y and z-axes by i j, and k respectively. In this coordinate system, the components of a vector u along x y , and z axes are denoted by ux uy , and uz . The vector u has the representation

u = uxi + uy j + uz k:
This notation does not lend itself to any abbreviation. Therefore, instead of denoting the coordinate axes by x

y z we will denote them by x1 x2 x3 . Also we will denote unit vectors

x2 and x3 axes by e1 e2, and e3 respectively. Naturally then components of a vector u along x1 x2 and x3 axes will be indicated by u1 u2 , and u3 respectively. Hence we can write u = u1e1 + u2e2 + u3e3
along x1

= uj ej :
Similarly,

(2.4.1)

v = v1e1 + v2 e2 + v3 e3 = vj ej :
7

The dot product u

v can simply be written as u v = u1v1 + u2v2 + u3v3 = uivi:


(2.4.2)

Since e1

e2 e3 are mutually orthogonal unit vectors, therefore, e1 e1 = e2 e2 = e3 e3 = 1 e1 e2 = e2 e3 = e1 e3 = 0: ei ej = ij :


(2.4.3)

These equations can be summarized as

Exercise. Using the index notation, write expressions for (1) the magnitude of a vector u, (2)

cos

being the angle between vectors u and v.

As another illustration of the use of the index notation, consider a line element with components

dx1 dx2 dx3 . The square of the length, ds, of the line element is given by
2 2 ds2 = dx2 1 + dx2 + dx3

= dxidxi = ij dxidxj :
Finally, we note that the differential of a function f (x1

x2 x3) can be written as

@f dx + @f dx + @f dx df = @x 1 @x2 2 @x3 3 1 @f dx : = @x i i
2.5 Permutation Symbol The permutation symbol, denoted by ijk , is dened by

ijk

=:

8 <

1 ;1 0

9 =

if i

j k form : an odd
not a
8

8 9 < an even =

permutation of

1 2 3:

(2.5.1)

That is,

123 132 111


We note that

= = =

231 321 211

= = =

312 213 133

=1 = ;1 = : : : = 0:

ijk

jki

kij

= ; jik = ; ikj = ; kji:

If e1 e2 e3 form a right-handed triad, then

e1 e2 = e3 e2 e3 = e1 etc:
which can be written as

ei ej =
Now, if u = ui ei

ijk ek :

(2.5.2)

v = vi ei, then u v = uiei vj ej = uivj ei ej = uivj ijk ek = ijk uivj ek :


(2.5.3) being the angle between

Exercise. Using the index notation write an expression for vectors u and v. Exercise. Show that

j sin j;

(1) (2)

ijk jki

=6 = 0:

ijk Aj Ak

The following useful identity, which can be veried by long-hand calculations should be memorized.

ijm klm

ik jl ; il jk :

(2.5.4)

Now by using this identity let us prove the vector identity

u (v w) = (u w)v ; (u v)w:
Proof: Let v

w = a. Then a =

ijk vi wj ek , and

u a = ijk uiaj ek aj = a ej = ilmvi wlem ej


=
ilm vi wl mj

ilj vi wl :

u (v w) =
= = (use (2:5:4)) = ( =

ijk ui( plj vpwl )ek ijk plj ui vpwl ek kij plj ui vpwl ek kp il ; kl ip )ui vp wl ek kp il ui vpwl ek ; kl ip ui vp wl ek

= ul wlvk ek ; upvpwk ek (use (2:4:2)) = (u w)v ; (u v)w:


Exercise. Show that (a) If ijk Tjk (b) ilm jlm (c) if Tij

= 0, then Tij = Tji,


ij , and

=2

= ;Tji, then Tij ai aj = 0.


2 3

We now write det Aij in the index notation.

A11 A12 A13 det Aij ] = det 4 A21 A22 A23 5 A31 A32 A33 = A11 (A22 A33 ; A32 A23 ) ; A21 (A12A33 ; A32 A13 ) + A31 (A12A23 ; A22 A13 )
= A11 ( 1jk Aj2Ak3 ) ; A21 (; 2jk Aj2Ak3) + A31 ( 3jk Aj2Ak3) = Ai1 =
ijk Aj 2 Ak3 ijk Ai1 Aj 2 Ak3 :
10

Example. Show that ijk Ail Ajm Akn

= (det A) lmn .

2.6 Manipulations with the Indicial Notations (a) Substitution If

ai = vim bm
and

(2.6.1)

bi = vim cm m to some other letter, say n and then the free index in (2.6.2) from i to m, so that bm = vmn cn:
Now (2.6.1) and (2.6.3) give

(2.6.2)

then, in order to substitute for bi s from (2.6.2) into (2.6.1) we rst change the dummy index from

(2.6.3)

ai = vim vmncn:

(2.6.4)

Note that (2.6.4) represents three equations each having the sum of nine terms on its right-hand side. (b) Multiplication If

p = am bm
and

(2.6.5)

q = cm dm
then

(2.6.6)

pq = am bm cndn:

(2.6.7)

11

It is important to note that pq

6= ambm cm dm. In fact the right-hand side of this expression is not

even dened in the summation convention and further it is obvious that 3 X pq 6= am bm cmdm: m=1 (c) Factoring If

Tij nj ; ni = 0
then, using the Kronecker delta, we can write

(2.6.8)

ni = ij nj
so that (2.6.8) becomes

(2.6.9)

Tij nj ;
Thus

ij nj

= 0: = 0:

(Tij ;
(d) Contraction

ij )nj

The operation of identifying two indices and so summing on them is known as contraction. For example, Tii is the contraction of Tij ,

Tii = T11 + T22 + T33


and Tijj is a contraction of Tijk ,

Tijj = Ti11 + Ti22 + Ti33 :


If

Tij = Ekk ij + 2 Eij


then

Tii = Ekk ii + 2 Eii


= (3 + 2 )Ekk :
12

Exercise. Given that Tij

1T E = Ekk ij + 2 Eij , W = 2 ij ij P = Tij Tij , show that

W = Eij Eij + 2 (Ekk )2


2.7 Translation and Rotation of Coordinate Axes

P = 4 2Eij Eij + (Ekk )2(4 + 3 2):

Consider two sets of rectangular Cartesian frames of reference O ; x1 x2 and O0 ; x01 x02 in a plane. If the frame of reference

O0 ; x01 x02 is obtained from O ; x1 x2 by a shift of the origin without a

change in orientation, then, the transformation is a translation.

x2) and (x01 x02 ) with respect to O ; x1 x2 and O0 ; x01 x02 respectively and (C1 C2 ) are the coordinates of 00 with respect to O ; x1 x2 , then
If a point P has coordinates (x1

x1 = x01 + C1 x2 = x02 + C2
or briey

xi = x0i + Ci i = 1 2:
If the origin remains xed, and the new axes Ox01 through an angle in the counter-clockwise direction, then

(2.7.1)

Ox02 are obtained by rotating Ox1 and Ox2

13

the transformation of axes is a rotation. Let the point relative to O ; x1 x2 and O ; x01 x02 respectively. Then,

have coordinates

(x1 x2) and (x01 x02 )

x1 = OA = OB ; CD
= OD cos ; PD sin = x01 cos ; x02 sin

x2 = AP = BD + CP
= x01 sin + x02 cos :
We can write (x01

x02) in terms of (x1 x2) as x01 = x1 cos + x2 sin x0


2

= ;x1 sin + x2 cos :

(2.7.2)

Using the index notation, the set of eqns. (2.7.2) can be written as

x0i = aij xj
where aij are elements of the matrix aij ];

i=1 2 j=1 2
sin cos

(2.7.3)

a11 a12 = cos a21 a22 ; sin


of arriving at (2.7.2). Let e01 along x1 and x2 -axes. Then

Before we generalize (2.7.1) and (2.7.2) to three dimensions we give below an alternate method

e02 denote unit vectors along x01 and x02-axes and e1 e2 unit vectors
;! OP = x1 e1 + x2 e2
= x01 e01 + x02 e02 :

Also

e01 = cos e1 + sin e2 e02 = ; sin e1 + cos e2 :


14

Therefore,

x01 = ;! OP e01
= (x1e1 + x2 e2 ) (cos e1 + sin e2 ) = x1 cos + x2 sin = ; x1 sin + x2 cos :
This latter approach can more easily be adopted to the 3-dimensional case. If the primed axes O ;

x02 = ;! OP e02

x01 x02 x03 are obtained from the unprimed axes O ; x1 x2 x3 just by a translation, then the coordinates of a point with respect to the two sets of axes are related by (2.7.1) wherein the index i ranges from 1 to 3. Now let us assume that the primed axes are obtained from the unprimed axes by a rotation only. Let us denote unit vectors along x1 x2 x3 by e1 e2 , and e3 respectively and those along x01 x02 x03 by e01 e02 and e03 respectively. If a1j = cosine of the angle between e01 and ej
then a11 write

a12 and a13 are the direction cosines of e01 with respect to the unprimed axes.

We can

e01 = a11 e1 + a12 e2 + a13 e3 = a1iei :


Similarly,

e02 = a2i ei e03 = a3iei :


Or

e0i = aij ej :
Note that the matrix aij is 3

(2.7.4)

3. Since

e0i e0j =
15

ij

(2.7.5)

therefore,

ij

= aik ek ajpep = aik ajpek ep = aik ajp


kp = aik ajk :
(2.7.6)

Equations (2.7.6) are equivalent to the following six equations.

2 2 a2 11 + a12 + a13 = 1 2 2 a2 21 + a22 + a23 = 1 2 2 a2 31 + a32 + a33 = 1


(2.7.7)

a11 a21 + a12 a22 + a13 a23 = 0 a21 a31 + a22 a32 + a23 a33 = 0 a31 a11 + a32 a12 + a33 a13 = 0:
three equations are equivalent to the statement that e01 we can write ei s in terms of e0i s. Since The rst three equations are equivalent to the statement that e01

e02 e03 are mutually orthogonal. Of course,

e02 and e03 are unit vectors; the last

aj1 = cosine of the angle between e1 and e0j


therefore,

e1 = aj1e0j or ei = ajie0j :
(2.7.8) must be identied as the inverse of the matrix aij :

(2.7.8)

From the point of view of the solution of a set of simultaneous linear equations, the matrix aji in

aij ];1 = aji] = aij ]T :


Here

(2.7.9)

aij ]T is the transpose of the matrix aij ]. A matrix aij ], which satises eqn. (2.7.9) is called aij ] in (2.7.4)

an orthogonal matrix. That is, the transpose of an orthogonal matrix equals its inverse. A transformation is said to be orthogonal if the associated matrix is orthogonal. The matrix dening a rotation of coordinate axes is orthogonal.

16

For an orthogonal matrix we have

a] a]T = 1:
Therefore

det( a] a]T ) = 1
or det a] det a]T

=1

or det a] det a] = 1 and thus

det a] = 1:
An orthogonal matrix whose determinant equals +1 is called proper orthogonal and the one whose determinant equals

;1 is called improper orthogonal.

A proper orthogonal matrix transforms a

right-handed triad of axes into a right-handed set of axes whereas an improper orthogonal matrix transforms a right-handed set of axes into a left-handed set of axes or vice-versa. Exercise: Consider a cube formed by the orthonormal vectors e01 of this cube equal to 1, show that det aij ] = 1. Consider a vector OP emanating from the origin O and ending at a point P . With respect to the primed and unprimed axes,

e02 and e03 . By setting the volume

OP = x01 e01 + x02 e02 + x03e03 = x0j e0j = xj ej :


Similarly,

x0i = OP e0i
= (xj ej ) (aik ek ) = xj aik = aij xj :
17

jk

Example: The components of a vector A with respect to unprimed axes are Ai

= (0 1 1). Con-

sider a set of primed coordinate axes obtained by rotating the unprimed axes through an angle of

30

about the x3 -axis (see Fig. ). What are the components, A0i , of this vector with respect to the

primed set of axes?

Solution:

e03 = e3 e01 = cos 30e1 + sin 30e2 e02 = ; sin 30e1 + cos 30e2:
Therefore

cos 30 sin 30 0 aij ] = e0i ej ] = 4 ; sin 30 cos 30 0 0 0 1


Now

3 5

A0i = aij Aj
can be written as
2

A01 cos 30 sin 30 0 0 sin 30 4 A0 5 4 5 4 5 4 1 = cos 30 5 : 2 = ; sin 30 cos 30 0 0 A3 0 0 1 1 1


Hence

32

A0i = (0:5 0:866 1):


18

Summarizing our discussion of the transformation of coordinate axes, we note that a general transformation from unprimed to primed axes combines both a translation and a rotation of the axes. This can be written as

x0i = aij xj + ci
of a vector A in the two sets of axes are related as

(2.7.10)

where aij is an orthogonal matrix and ci is a constant. Under this transformation, the components

A0i = aij Aj :
2.8 Tensors
2.8.1 A Linear Transformation

(2.7.11)

Let T be a transformation from a vector space into the same vector space. That is, for any vector

u, Tu is also a vector of the same dimensions as u. Then T is linear if and only if T( u + w) = Tu + Tw for every real
[NOTE: f is a linear function of x if and only if f (x) = and

(2.8.1.1)

x where

is a real number. For example

f (x) = 2x + 3 is not a linear function of x even though it is often referred to as such.]


A linear transformation from a vector space into another vector space is also called a second-order tensor.
2.8.2 Tensor Product Between Two Vectors

The tensor product (or the dyadic product)

between two vectors a and b is dened as (2.8.2.1)

(a b)c = (b c)a for every vector c:


Thus (a

b) transforms a vector c into a vector parallel to a. Since it transforms a vector into a a b 6= b a:


(2.8.2.2)

vector and obeys (2.8.1.1), it is a linear transformation. Note that

19

2.8.3 Components of a Second-Order Tensor

^2 e ^3 be a set of orthonormal basis vectors (i.e. e ^1 ^ ^3 are mutually orthogonal unit e e2 e vectors). For any vector a, ^i : a = aie
Let b = Ta. Then (2.8.3.1)

^1 Let e

^j ) = aj (Te ^j ) b = T(aj e
or

(2.8.3.2)

^i = aj (Te ^ j ): bie ^k , we obtain Taking the inner product of both sides of this eqn. with e ^k aj (Te ^j ) = aj e ^k (Te ^j ) = aj Tkj bk = e
where

(2.8.3.3)

(2.8.3.4)

^k (Te ^j ) Tkj = e ^i . is called the component of T with respect to the basis e


For computation purposes, eqn. (2.8.3.4) is written as
8 9 < b1 = :

(2.8.3.5)

b2 b3

T11 T12 T13 = 4 T21 T22 T23 T31 T32 T33

38 < 5 :

a1 a2 a3

9 =

(2.8.3.6)

second-order tensor T.

Analogous to the representation (2.8.3.1) for vector a, we have the following representation for

^i e ^j : T = Tij e

(2.8.3.7)

Because of (2.8.2.2), Tij need not equal Tji . In order to see that (2.8.3.7) is equivalent to (2.8.3.5),

20

we evaluate Ta.

^i e ^j )(ak e ^k ) b = Ta = (Tij e ^j )^ = Tij ak (^ ei e ek ^i(^ ^k ) = Tij ak e ej e ^i jk = Tij aj e ^i = Tij ak e


which is equivalent to bi

(2.8.3.8)

^i. Let It is clear from (2.8.3.7) that the components Tij of T depend upon the choice of basis e ^0i = Qij e ^j e
(2.8.3.9)

= Tij aj or (2.8.3.4).

where Q is an orthogonal matrix (i.e.

QQT = 1). Then


(2.8.3.10)

0e 0 (Q e ^0i e ^0j = Tij ^l ) = T = Tij ik ^k ) (Qjl e 0 Q Q (^ ^l ) = Tkl e ^k e ^l : = Tij ik jl ek e


Hence

0Q Q Tkl = Tij ik jl
and in matrix notation,

(2.8.3.11)

T ] = Q]T T 0] Q]
and since Q is orthogonal,

(2.8.3.12)

T 0] = Q] T ] Q]T :
The transpose TT of a second-order tenseor T is dened by

(2.8.3.13)

a (TT b) = b (Ta) for every vector a and b:


The components of T and TT are related by

(2.8.3.14)

(T T )ij = Tji
21

(2.8.3.15)

2.8.4 Tensors of Higher-Order

A third-order tensor is a linear transformation from the space of second-order tensors into vectors or vectors into second-order tensors, and can be represented as

^i e ^j e ^k : A = Aijk e
follows.

(2.8.4.1)

Under a change of basis (2.8.3.9), the transformation rule for its components can be derived as

^0i e ^0j e0k A = A0ijk e ^l Qjme ^m Qkne ^kn = A0ijk Qil e ^l e ^m e ^n = A0ijk Qil QjmQkne ^l e ^m e ^n : = Almn e
Thus

(2.8.4.2)

Almn = Qil QjmQknA0ijk


or

(2.8.4.3)

A0ijk = Qil QjmQknAlmn:


second-order tensors, and has the representation

(2.8.4.4)

A fourth-order tensor is a linear transformation from the space of second-order tensors to

^i e ^j e ^k e ^l : C = Cijkle
Under the transformation (2.8.3.9) its components will transform as follows.

(2.8.4.5)

0 =Q Q Q Q C : Cijkl ip jq kr ls pqrs

(2.8.4.6)

3 Kinematics
3.1 Description of Motion of a Continuum. Let us suppose that a body, at time t

= t0, occupies a region of the physical space. The position of a particle at this time can be described by its coordinates Xi with respect to a xed rectangular
22

Cartesian coordinate system.

Let the body undergo a motion and point P move to P 0 whose coordinates with respect to the same xed axes are xi . Then an equation of the form

xi = xi(X1 X2 X3 t)
describes the path of the particle which at

(3.1.1)

t = t0

is located at

Xi.

In eqn. (3.1.1) the triplet

(X1 X2 X3) serves to identify different particles of the body and is known as reference coordinates. The triplet (x1 x2 x3 ) gives the present position of the particle which at time t = t0 was at the place Xi . Note that for a specic particle eqn. (3.1.1) denes the path line (or trajectory) of
the particle. Of course

Xi = xi(X1 X2 X3 t0 )
which merely veries the fact that the particle under consideration occupied the place Xi at t = t0 . Example: Consider the motion

x1 = X1 + 0 2tX2 x2 = X2 x3 = X3
23

(X1 X2 X3 ) gives the position of a particle at t = 0. Sketch the conguration at time t = 2 for the body which at t = 0 has the shape of a cube of unit sides with one corner at the
where origin. Solution: For the particle which at t = 0 was at the origin,

xi = 0 for all t:
Thus this particle stays at the origin at all times.

Similarly, the particle which at time t = 0 was at the position (X1 That is, particles on line OA do not move. A particle (X1

0 0) will move to xi = X1

1i .

1 0) on line CB will occupy, at time t = 2, the position

xi = (X1 + 0 2(2) (1)) 1i + (1) 2i:


Thus every particle on line CB is displaced horizontally to the right through a distance (0

2) (2) =

0 4.
A particle (0

X2 0) on line OC moves to xi = (0 + 0 2(2)X2) 1i + X2


2i

so that every particle on the line

OC moves horizontally to the right through a distance linearly

proportional to its height, that is, it remains a straight line. A similar situation prevails for the line

BA.
24

Thus at time shown.

t = 2, the side view of the cube changes from a square to a parallelogram as

The motion given in this example is known as simple shearing motion. 3.2 Referential and Spatial Descriptions When a continuum is in motion, quantities (such as temperature , velocity v) that are associated with specic particles change with time. There are two ways to describe these changes. (i) Following the particles, that is, we express , vi as functions of the coordinates of a particle in a xed reference conguration and time t. Said differently, we express

= (X1 X2 X3 t)

vi = vi(X1 X2 X3 t):
Such a description is known as Lagrangian or referential or material description. (ii) Observing the changes at xed locations, that is, we express

vi etc. as functions of xi and

t. Thus
= (x1 x2 x3 t)

vi = vi(x1 x2 x3 t):
Such a description is known as spatial or Eulerian. We note that in this description, what is described (or measured) is the change of quantities at a xed point in space (not a specic material particle) as a function of time. The same spatial position is occupied by different particles at different times. Therefore, the spatial description does not provide direct information regarding the changes in the values of a quantity associated with a material particle as it moves about in space. The referential and spatial descriptions are, of course, related by the motion. That is, if the motion is known then, one description can be obtained from the other as illustrated by the following example.
25

Example: Given the motion of a body to be

xi = Xi + 0 2tX2 1i:
For the temperature eld given by

(i)

= 2x1 + (x2)2
(a) nd the material description of temperature, and (b) the rate of change of temperature of a particle which at time t = 0 was at the place (0 Solution (a) By substituting (i) into (ii), we obtain

(ii)

1 0).

= 2x1 + (x2 )2 = 2(X1 + 0 2tX2 ) + (X2)2 = 2X1 + (X2 + 0 4t)X2:


(b) The temperature of the desired particle at different times is given by (iii)

d = 0 4: dt

= (1 + 0 4t):

Note that even though the temperature is independent of time in the spatial description, a particle experiences a change in temperature as it moves from one spatial position to another; this becomes clear from eqn. (iii). Whereas spatial description is often used in uid mechanics, referential description is employed in solid mechanics and in formulating laws of mechanics. 3.3 Displacement Vector By denition, the displacement vector ui of a particle is the difference between its position vectors at time t and at time t = t0 (or 0).

ui = xi ; Xi:
26

(3.3.1)

In the Lagrangian description, the displacement ui is specied as a function of example, consider the motion

Xi and t.

For

x1 = X1t2 + 2X2t + X1 x2 = 2X1t2 + X2t + X2 x3 = 1 X t + X3 : 2 3


The corresponding components of the displacement are given by (3.3.2)

u1 = X1t2 + 2X2t u2 = 2X1t2 + X2t u3 = 1 2 X3t:


In the Eulerian description, ui will be expressed as a function of xi and t. Solving (3.3.2) for Xi in terms of xi and t and substituting that in (3.3.1) we obtain

x1 (1 + t) ; 2x2 t u1 = x1 ; X1 = x1 ; ; 3t3 + t2 + t + 1 2 2 2 (1 + t ) ; 2x1 t u2 = x2 ; X2 = x2 ; x ;3t3 + t2 + t + 1 x3 : u3 = x3 ; X3 = x3 ; t2+ 2


3.4 Restrictions on Continuous Deformation of a Deformable Body. For t = 1, eqns. (3.3.2) give

x1 = 2(X1 + X2 ) x2 = 2(X1 + X2 ) x3 = 3 2 X3:


Thus points (a

;a 0) in the reference conguration move to (0 0 0) at time t = 1. This implies


= 1).
This amounts to the collision of

that distinct particles which occupy different places in the reference conguration are deformed into the same place in the present conguration (at time t

various material particles. Even though in particle mechanics collisions are allowed, in continuum
27

mechanics, such a possibility is assumed to be ruled out to start with. Thus in continuum mechanics it is assumed that different material particles always occupy distinct places. This is equivalent to the requirement that

xi = xi(X1 X2 X3 t)
be a one-to-one mapping of R onto the present conguration. Since in Continuum Mechanics we will need to differentiate functions with respect to xi or mapping

Xi, we will henceforth assume that the

xi = xi(X1 X2 X3 t)
is continuously differentiable and has a continuously differentiable inverse given by

Xi = Xi(x1 x2 x3 t):
This is so if and only if the Jacobian J dened by

@x1 @x1 @X2 @X3 @x2 @x2 @X2 @X3 @x3 @x3 @X2 @X3 is non-zero for all points in R and for every value of t. Since @x1 6 @X1 6 6 @x2 J = det 6 6 6 @X1 6 4 @x3 @X1 J (X1 X2 X3 0) = 1

3 7 7 7 7 7 7 7 5

(3.4.1)

and

is a continuous function of t, therefore,

must be positive for every t. Using the index

notation, we can write J as

@xi = det J = det @X j

ij +

@ui : @Xj

(3.4.2)

In summary, we can conclude from the preceding discussion that a necessary and sufcient condition for a continuous deformation to be physically admissible is that the Jacobian J be greater than zero.
28

A displacement eld satisfying the condition simply admissible.

J > 0 is said to be proper and admissible, or u2 u3)

Thus, for an admissible deformation of a medium the displacement components (u1 must satisfy J

> 0. For example, a piece of rubber can not be subjected to displacement components u1 = ;2X1 u2 = 0 u3 = 0 since then J = ;1. This displacement is called a reection about the (X2 X3 ) plane, since the point (x1 x2 x3 ) is the image of the point (X1 X2 X3 ) in a mirror that lies in the plane x1 = 0.
Exercise: Determine whether or not

u1 = k(X2 ; X1) u2 = k(X1 ; X2) u3 = kX1 X3


where k is a constant, are possible displacement components for a continuous medium. 3.5 Material Derivative The time rate of change of a quantity such as temperature or velocity of a material particle is known as a material derivative, and is denoted by a superimposed dot or by D=Dt. (i) When referential or Lagrangian description of a quantity is used, i.e.,

= (X1 X2 X3 t)
then

_=D =@ Dt @t Xi;xed:
(ii) When spatial or Eulerian description of a quantity is used, i.e.,

(3.5.1)

= (x1 x2 x3 t)
where

xi = xi (X1 X2 X3 t)
then

_=D =@ +@ Dt @t xi;xed @xj


29

@xj @t

Xi ;xed

(3.5.2)

In rectangular Cartesian coordinates

j vj = @x @t

is the j th component of the velocity of a material particle. Therefore, in rectangular Cartesian coordinates,

Xi ;xed

(3.5.3)

_ = @ + @ vj : @t @xj
Note that in (3.5.4) is given in the spatial description. Example: Given the motion

(3.5.4)

xi = Xi(1 + t) t 0
nd the spatial description of the velocity eld. Solution

xi = Xi(1 + t)

i vi = x _ i = Xi = 1 x + t:

Example: Given

2 = 2(x2 1 + x2 ) where xi = Xi (1 + t):

t = 1, the rate of change of temperature of the material particle which in the reference conguration was at (1 1 1).
Find, at Solution (i)

2 = 2(x2 1 + x2 )

= 2 X12(1 + t)2 + X22(1 + t)2] _ = 2 2X12(1 + t) + 2X22(1 + t)] (ii)


) _ at t = 1 for the material particle (1 1 1) = 16: _ = @ + @ @xj

@t @xj @t = 0 + 4x1 x1 + 4x2 x2 : 1+t 1+t


30

At t = 1, for the material particle (1

1 1)

xi = 2( i1 + i2 + i3 ) 2 2 _ = 4(2) + 4(2) = 16: 1+1 1+1

Exercise: The motion of a continuous medium is dened by the equations

x3 = X3

x1 = 1 (X1 + X2 )et + 1 (X1 ; X2)e;t x2 = 1 (X1 + X2)et ; 1 (X1 ; X2 )e;t 2 2 2 2


0 t < constant:

(a) Express the velocity components in terms of referential coordinates and time. (b) Express the velocity components in terms of spatial coordinates and time. Exercise: Given the motion of a body to be xi temperature eld by

= x1 + x2 , nd _ for the particle which currently is at the place (1 1 1).

= (X1 + ktX2 )

i1

+ X2

i2

+ X3

i3 , and the

3.6 Finding Acceleration of a Particle from a Given Velocity Field. The acceleration of a material particle is the rate of change of its velocity. If the motion of a continuum is given by

xi = xi (X1 X2 X3 t) with xi (X1 X2 X3 0) = Xi


then the velocity v, at time t, of a particle Xj is given by

i vi = @x @t i ai = @v @t
Thus, if the material description of velocity

and the acceleration a, at time t, of a particle Xj is given by

Xj ;xed

=x _i

Xj ;xed

=v _i:

v(X1 X2 X3 t) is known, then the acceleravi =

tion is computed simply by taking the partial derivative with respect to time of the function

v(X1 X2 X3 t).

On the other hand, if only the spatial description of velocity (i.e.

31

vi(x1 x2 x3 t)) is known, then the computation of acceleration involves the use of eqn. (3.5.2).
That is,

= @vi + @vi @xk @t xj ;xed @xk @t Xj ;xed Xj ;xed @vi v : i = @v + @t xj ;xed @xk k @vi is called convective acceleration. When the motion is The part of acceleration given by vk @xk only along x1 -axis, i.e. v2 = v3 = 0 and v1 = v1 (x1 t), then @v1 : 1 a1 = @v + v 1 @t @x1 Example: A uid rotates as a rigid body with a constant angular velocity ! e3 .
(a) Write out explicitly the components of the velocity of a material particle in the Lagrangian and Eulerian descriptions of motion. (b) Compute the acceleration eld in the Eulerian description. Solution:

i ai = @v @t

(a) Recall that

v = ! r:
32

Thus

vi = "ijk !j xk = "ijk ! j3xk


= !"i3k xk
is the Eulerian description of velocity. To convert these expressions into the Lagrangian description, let the uid particle which presently is at place P be at place Q at t = 0. Then, referring to Fig. 3.6.1,

; = !t
Since

X2 + !t: = + !t = tan;1 X 1

x1 = OP cos
= OQ cos
q 2 v1 = !"132x2 = ;! X12 + X22 sin(tan;1 X X1 + !t) q X2 + !t) : v2 = !"231x1 = ! X12 + X22 cos(tan;1 X 1

= X12 + X22 cos

33

(b)

i @vi ai = @v @t + @x vj j

= 0 + !"i3k kj vj = !"i3j (!"j3mxm )

= !2"i3j "3mj xm = ! 2(
)
i3 3m ; im 33 )xm

ai = !2( i3 x3 ; xi) : a1 = ; !2x1 a2 = ; !2x2 a3 = 0 :


Exercise: Given

e1 + x2 e2 v = x1x 2 + x2
1 2
currently is at the place (1 Exercise: For the motion

2 = 2(x2 1 + x2 )

determine the acceleration and the rate of change of temperature of the material particle which

1).

xi = Xi + sin( t) sin( X1)

i2

nd the velocity and acceleration in referential and spatial descriptions. 3.7 Deformation Gradient As pointed out earlier, in continuum mechanics, two different material particles always occupy two distinct places. Consider two material particles P (Xi) and Q(Xi + dXi ).

34

X2 X3) used to identify particles in the reference conguration. Let points P and Q move to P 0 and Q0 respectively so that the vector PQ is deformed into

Here (Xi ) stands for the triplet (X1 the vector P0 Q0 . The vector

P0Q0 need not, and in general will not, equal PQ. This means that the length and direction of P0 Q0 may be different from that of PQ. Given PQ or P0 Q0 and the
Let the motion of the body be given by

motion, our problem is to nd the other vector.

xi = xi (X1 X2 X3 t):
Then

(3.7.1)

P0Q0 = xi (X1 + dX1 X2 + dX2 X3 + dX3 t) ; xi (X1 X2 X3 t)]ei:


Using Taylors theorem for series expansion, we get

(3.7.2)

@xi dX + @xi dX + @xi dX e + O(jdX j2) P0Q0 = @X 1 @X 2 @X 3 i @xi dX e + O(jdX j2): = @X j i


j
1 2 3

(3.7.3) (3.7.4)

If points P and Q are close together in the reference conguration, then higher order terms can be neglected as compared to terms linear in dXi . Using this approximation, we get

@xi dX e : P0Q0 = @X j i
j P
Here the sufx P reminds us that

(3.7.5)

@xi is evaluated at the point P . In section 2.9, it was mentioned @Xj that a comma followed by an index i will be used to indicate partial differentiation with respect to xi . Now if we were to use that notation, then it will not be clear whether the partial differentiation
35

is with respect to xi or Xi . To clarify the situation we will, henceforth, use upper case latin indices for X . That is, the triplet (X1 be written as

X2 X3) will be denoted by XA instead of Xi and eqn. (3.7.5) will


(3.7.6)

@xi dX e = F dX e : P0Q0 = @X A i iA P A i
A P
In component form,

(P 0Q0 )j = P0Q0 ej = FiAjP dXAei ej = FjA P dXA = FjA P (PQ)A :


(3.7.7) Thus FjA relates the components of vector PQ in the reference conguration to the components of the vector P0 Q0 into which PQ is deformed. Since in Continuum Mechanics we assume that

J = det jFiAj > 0


therefore, FjA is an invertible matrix. This implies that once the motion (3.7.1) is known, we can nd P0 Q0 from PQ and vice-a-versa. Since the motion (3.7.1) gives how the body deforms and FiA is the gradient of the motion, FiA

is called deformation gradient. FiA relates a vector PQ in the reference conguration to the vector

P0Q0 (in the present conguration) into which PQ is deformed. In terms of the displacement u,
FiA can be written as follows. ui = xi ; XA iA @ui = @xi ; @XA @XA iA FiA = xi A = iA + ui A :

(3.7.8)

The gradient ui A of the displacement u is known as the displacement gradient. Example: The deformation of a body is given by

u1 = (3X12 + X2) u2 = (2X22 + X3) u3 = (4X32 + X1 ): 1 1 1 passing through the point (1 1 1) in Compute the vector into which the vector 10;2 3 3 3
the reference conguration is deformed.
36

Solution Here the point

is

;2 ;2 ;2 (1 1 1) and the vector PQ has components 10 10 10 3 3 3


2 3 5

. The

deformation gradient at any point is given by

1 + 6X1 1 0 4 0 1 + 4X2 1 FiA] = 1 0 1 + 8X3 The deformation gradient evaluated at point P is 2 3 7 1 0 FiA] P = 4 0 5 1 5 : 1 0 9
Therefore, the components of the vector P0 Q0 are given by 2 3 10;2
2 6 6 4

(P 0Q0)1 (P 0Q0)2 (P 0Q0)3

7 1 0 7 6 7=6 0 5 1 5 4 1 0 9

6 6 76 76 56 6 4

PQ in magnitude. The ratio of the length of P0Q0 to that of PQ is called stretch in the direction of PQ.
It is emphasized that in going from (3.7.4) to (3.7.5) we assumed that the length of the vector

Note that, in the preceding example, the vector P0 Q0 is neither parallel to PQ nor is it equal to

3 10;2 3 10;2 3

7 7 7 7 7 7 5

= 103

;2 6
6 4

8 7 6 7 5: 10

PQ is innitesimal. However, no assumption was made as to the magnitude1 of FiA . Thus (3.7.5)
is valid no matter how large or small the components of deformation gradient are. Said differently (3.7.5) is valid both for small and large deformations so long as we study the deformation of innitesimal vectors passing through the point P . Thus FiA describes the deformation of material particles in an innitesimal neighborhood of P . A special case in which (3.7.5) follows exactly from (3.7.4) is that when FiA is constant. That is, each of the nine quantities @xi =@XA is a constant. A deformation for which FiA is a constant is called a homogeneous deformation. Example: Given the following displacement components

1 The

magnitude of FiA is dened as

p1=2FiAFiA.

u1 = 0 1X22 u2 = u3 = 0:
37

a) Is this deformation possible in a continuously deformable body? Prove your answer. b) Find the deformed vectors of the material vectors 0:01e, and 0:015e2 which pass through the point P

(1 1 0) in the reference conguration. 1 0) in the X1 and X2 directions. (1 1 0) that were parallel

c) Determine the stretches at the point (1

d) Determine the change in angle between lines through the point P to X1 and X2 axes. Solution: a)

1 0:2X2 0 FiA] = 4 0 1 0 5 0 0 1 det FiA] = 1 > 0:


Therefore, the given deformation is possible in a continuously deformable body. b) The material vector 0 given by
2 4

01e1 through the material point (1 1 0) is deformed into the vector 1 0:2(1) 0 0:01 0:01 0 1 0 54 0 5 = 4 0 0 0 1 0 0 1 0:2(1) 0 0 1 0 0 0 1
32 54 3 2 32 3 2 3 5

and the material vector 0:015e2 through the material point (1


2 4

1 0) is deformed into 3 0 0:003 0:015 5 = 4 0:015 5 : 0 0

c) Stretch at P (1

d)

:01 = 1. 1 0) in X1-direction = 0 0:01 p 0:0032 + :0152 = 1:02. Stretch at P (1 1 0) in X2 -direction = 0:015 Angle between the vectors into which vectors 0:01e1 and 0:015e2 through the point (1 1 0) (0:01)(0:003 + 0 + 0) = 78:7 . are deformed = cos;1 (0:01)(0:0032 + 0:0152)1=2 Change in angle = 11:3 .
38

Example: Given the following displacement components

u1 = 2X12 + X1 X2 u2 = X22 u3 = 0
and that for points in the body, X1

0 X2

0.

a) Find the vector in the reference conguration that ends up parallel to x1 -axis through the point

(1 0 0) in the current conguration.


b) Find stretch of a line element that ends up parallel to x1 -axis through the point (1 present conguration. Solution: For the given displacement components,

0 0) in the

x1 = X1 + 2X12 + X1X2 x2 = X2 + X22 x3 = X3:


Therefore the undeformed position of point (1

0 0) is obtained by solving

1 = X1 + 2X12 + X1 X2 0 = X2 + X22 0 = X3:


The solution of these equations which satises X1 currently is at P 0 (1

0 is (1=2 0 0). The material particle which

0 0) was at P (1=2 0 0) in the reference conguration. 2 3 1 + 4X1 + X2 X1 0 0 1 + 2X2 0 5 FiA] = 4 0 0 1 2 3 3 0:5 0 4 FiA P = 0 1 0 5 0 0 1 2 3 2 32 3 1 3 0:5 0 dX1 ) 4 0 5 = 4 0 1 0 5 4 dX2 5 0 0 0 1 dX3 where (dX1 dX2 dX3) is the vector in the reference conguration that ends up into a (1 0 0) through the point (1 0 0) in the current conguration. 3 2 2 32 3 2 3 1 ;0:5 0 dX1 1 1=3 4 dX2 5 = 1=3 4 0 3 0 54 0 5 = 4 0 5: dX3 0 0 3 0 0
39

vector

Thus a vector parallel to x1 -axis through (1=2 vector parallel to x1 -axis through (1

0 0) in the reference conguration ends up into a

0 0) in the current conguration.

b) Stretch along the desired line = p Exercise: Given the displacement eld

12 + 0 + 0 = 3. (1=3)2 + 0 + 0

u1 = 10;2(2X1 + X22) u2 = 10;2(X12 ; X22) u3 = 0:


Find the stretches and the change of angle for the material lines emanate from the material particle (1

(0:1 0 0) and (0 0:1 0) that

;1 0).

Exercise: The displacement components for a body are

u1 = 2X1 + X2 u2 = X3 u3 = X3 ; X2 :
a) Verify that this displacement vector is possible for a continuously deformed body. b) Is this deformation homogeneous? c) Determine the stretch in the direction (1=3 ence conguration. d) Determine the direction cosines of the line element in the reference conguration that ends up in the x3 -direction at the place (1

1=3 1=3) through the point (1 1 1) in the refer-

1 0).

e) Determine the change in angle between the lines through the point (1

1 1) (in the reference conguration) whose directions in the reference conguration are 1 0 0 and 1=3 1=3 1=3.

3.8 Strain Tensors As seen in the previous section, during the motion of a deformable continuum, material lines originating from a material point are rotated and stretched. Whenever the material lines emanating from a material point are stretched and/or the angle between two different material lines passing through a material point changes, the body is said to be strained or deformed. We have seen in the previous section that the deformation gradient is a measure of the stretch and rotation of various
40

material vectors. In this section we will introduce two more measures of deformation which, of course, will be related to the deformation gradient. reference conguration. Let PQ be deformed into P0 Q0 and PR into P0 R0 . Then Consider two material vectors PQ and PR originating from the material point P (XA) in the

(P 0Q0 )j = FjA (PQ)A (P 0R0 )


the material point P . Therefore,

= FjA (PR)A:
P

Henceforth we will drop the sufx jP to shorten the notation. Of course, FiA is to be evaluated at

P0Q0 P0R0 = FjAFjB (PQ)A(PR)B


= (PQ)A CAB (PR)B
where (3.8.1)

CAB = FjAFjB or C] = F]T F]:


Note that CAB

(3.8.2)

= CBA , that is the matrix C] is symmetric. To obtain a physical interpretation of various components of C], let

PQ = 10;2(1 0 0) PR = 10;2(1 0 0):


41

Then (3.8.1) gives

P0Q0 P0Q0 = 10;4C11


so that

jP0Q0 j = 10;2 C11 or jP0Q0j=jPQj = C11:


X2 and X3 axes respectively. Now take

(3.8.3)

Thus C11 is a measure of the stretch along X1 -axis. Similarly C22 and C33 measure stretches along

PQ = 10;2(1 0 0) PR = 10;2(0 1 0):


Then (3.8.1) gives

P0Q0 P0R0 = 10;4C12 : p p Since jP0Q0j = 10;2 C11 jP0R0j = 10;2 C22 0 0 0 0 p : (3.8.4) therefore P0Q0 P0R0 = p C12 jP Q j jP R j C11 C22 The left-hand side of (3.8.4) equals cosine of the angle between P0 Q0 and P0 R0 . Thus C12 provides
a measure of the change in angle between two material lines passing through the point P that in the reference conguration were parallel to X1 and X2 -axes. Similarly C23 measures the change in angle at the material point P between two material lines that in the reference conguration were parallel to X2 and X3 axes. In terms of displacement components CAB can be written as follows.

CAB = FiA FiB = (


=

iA + ui A)( iB + ui B )
(3.8.5)

AB + uB A + uA B + ui Aui B :

CAB is called the right Cauchy-Green tensor. The tensor EAB = 1=2(CAB ;
AB )
(3.8.6)

= 1=2(uA B + uB A + ui Aui B )
42

is known as the Green-St. Venant strain tensor. We note from (3.8.1) that

P0Q0 P0R0 ; PQ PR = (PQ)A(CAB ;


so that

AB )(PR)B
(3.8.7)

= 2(PQ)AEAB (PR)B

EAB measures the change in lengths of various material line elements and the change in

angles between different material lines emanating in the reference conguration from the same material point. It follows from eqn. (3.8.7) that

P0Q0 P0R0 ; PQ PR = 2M E N A AB B jPQj jPRj jP0Q0 j2 ; jPQj2 = 2M E M A AB B jPQj2 where M and N are unit vectors parallel to PQ and PR respectively.
Since (PQ)A

(3.8.8)

= (F ;1)Ai(P 0Q0 )i, therefore

PQ PR = (F ;1)Ai(F ;1)Bj (P 0Q0 )i(P 0R0 )j


= (B ;1 )ij (P 0Q0 )i(P 0R0 )j
where (3.8.9)

Bij = FiAFjA B] = F] FT ]
is the left Cauchy-Green tensor. The tensor

(3.8.10)

ij

1 ( ; (B ;1) ) =2 ij ij =1 2 (ui j + uj i ; uk iuk j )

(3.8.11)

is called the Almansi-Hamel tensor. It follows from (3.8.9) and (3.8.11) that

P0Q0 P0R0 ; PQ PR = 2 m n (3.8.12) ij i j jP0Q0j jP0R0j where m and n are unit vectors parallel to P0 Q0 and P0 R0 in the present conguration. It is evident
from eqn. (3.8.12) that ij also measures changes in lengths of line elements and changes in angles between different line elements in the present conguration.
43

In the referential description of motion,

EAB is used as a measure of strain.

However, ij is

used to measure strain in the spatial description of motion. Note that each one of these two tensors vanishes when there is no deformation. Exercise: By taking PQ = 10;2 (1

0 0) PR = 10;2(1 0 0), and using (3.8.7), prove that


p

jP0Q0 j =
Exercise: By setting PQ = 10;2 (1

1 + 2E11 10;2:

Hence nd the engineering strain (elongation/original length) along X1 -axis. of the previous exercise, obtain an expression for the cosine of the angle between P0 Q0 and P0 R0 .

0 0) and PR = 10;2(0 1 0) into (3.8.7), and using the result

3.9 Principal Strains Having learned how to nd the deformation of various material vectors emanating from a material point P , we now investigate which of these lines is stretched the most. Let

PQ = 10;3(N1 N2 N3)
where

(3.9.1) That is,

N = (N1 N2 N3) is a unit vector along PQ. Our aim is to nd N such that the stretch at P along N is maximum or minimum. The vector P0Q0 into which PQ is deformed is given by
(P 0Q0 )j = FjA(PQ)A = 10;3FjANA:
Therefore, (3.9.2)

(N1 N2 N3 ) are components of a unit vector along PQ.

jP0Q0j2 = F N F N = C N N : (3.9.3) jPQj2 jA A jB B AB A B Thus the problem reduces to nding a unit vector N such that CAB NA NB is maximum. By using the method of Lagrange multipliers we need to nd N such that
CAB NANB ; (NANA ; 1)
takes on extreme values for all N. In (3.9.4) (3.9.4) is a Lagrange multiplier. Such an N is given by

@ C N N ; (N N ; 1)] = 0 A A @NI AB A B
44

which is equivalent to

CAB NB ; NA = 0
or

(CAB ;

AB )NB

= 0:

(3.9.5)

These are three linear homogeneous equations in N1 tions exists if and only if

N2 N3. A non-trivial solution of these equa-

det CAB ;
or

AB ] = 0

3 ; I 2 + II
where

; III = 0

(3.9.6)

I = CAA II = 1=2 ;CAB CBA + CAACBB ] = (det C])(C ;1)AA III = det C]: I II III are called principal invariants of C]. Equation (3.9.6) is cubic in and will have three roots. Since the matrix C] is symmetric and positive denite, all three roots of (3.9.6) are positive.
nd the corresponding N from (3.9.5). That is, once the three roots of (3.9.6) have been found, we use (3.9.5) to nd the directions N(1) For example, N(1) is obtained by solving Let us denote the three roots of (3.9.6) by 2 1

2 2

2 .2 For each one of these roots of (3.9.6) we 3

N(2) N(3) along which stretches assume extreme values.


(3.9.7)

(CAB ;
2

(1) 2 1 AB )NB

= 0 N1(1) N1(1) + N2(1) N2(1) + N3(1) N3(1) = 1:

2 3 should not be confused with the components of a vector.

45

Having found N(1) , we use (3.9.3) to nd the stretch in this direction. Stretch along N(1)

(1) N (1) = CAB NA B

q q

2 AB N (1) N (1) 1 A B

2 N (1) N (1) 1 A A

= 1:
Thus the three roots of (3.9.6) are squares of extreme values of stretches at the point P . Let us now assume that 2 1 as

6=

2 2

6=

2 and nd the angle between directions N(1) and N(2) . Rewriting (3.9.7) 3

(1) CAB NB =
we obtain

(1) 2 1 AB NB

(1) = 2 1 NA

(2) (1) (2) (1) NA CAB NB = 2 1 NA NA :


Similarly

(3.9.8)

(1) (2) (1) (2) NA CAB NB = 2 2 NA NA :


Since CAB

(3.9.9)

= CBA, therefore, the left-hand sides of (3.9.8) and (3.9.9) are equal. Thus
2 (1) (2) ( 2 1 ; 2 )NA NA = 0:

Since 2 1

6=

2 by assumption, 2 (1) (2) NA NA = 0:


(3.9.10)

N(1) and N(2) are perpendicular to each other. By using the same argument for N(2) and (1) N(2) and N(3) are mutually perpendicular N(3) , we conclude that whenever 2 6 2 6 2 1= 2= 3, N
Thus to each other. Let us now nd the change in the angle between the lines N(1) and N(2) during the deformation.

46

If is the angle between the deformed positions of N(1) and N(2) , then by (3.8.1) and (3.8.4)

(1) (2) C FGNF NG cos = (1) (1) 1=2 (2) (2) 1=2 (CAB NA NB ) (CDE ND NE )

=
)

= 90 :

(1) (2) 2 1 AB NB NA 1 2

(1) (2) = 1NA NA = 0: 2


(3.9.11)

same holds true for directions N(2) and N(3) . Because of this property, the directions N(1) and N(3) are called principal axes of stretch and 1 suitable choices of N(1)

That is, the angle between directions N(1) and N(2) does not change during the deformation. The

N(2)

2 2 2

3 are called principal stretches.

Whenever any two or all roots of eqn. (3.9.6) are equal, equations (3.9.10) and (3.9.11) hold for

2 , then N(3) is uniquely determined but 3 N(1) and N(2) can be taken as any two directions in a plane perpendicular to N(3) . When 2 1 = 2 = 2 , then any three linearly independent directions which need not be mutually perpendicular 2 3 to each other can be taken as N(1) N(2) , and N(3) . 2 1

N(2) and N(3) . If

6=

Recalling that

2EAB = CAB ;
we obtain

AB

(1) (1) 2EAB NB = CAB NB ;

(1) AB NB
(3.9.12)

(1) =( 2 1 ; 1)NA :
To arrive at this equation, we used (3.9.7). Thus N(1)

N(2) and N(3) are the directions along which


(3.9.13)

EAB NANB takes on extreme values


2 2 ( 2 1 ; 1)=2 ( 2 ; 1)=2 ( 3 ; 1)=2
respectively. If we dene strain in the direction PQ as

(jP0Q0j2 ; jPQj2)=jPQj2
47

(3.9.14)

instead of the engineering strain

(jP0Q0j ; jPQj)=jPQj

(3.9.15)

we see from (3.8.7) that EAB NA NB gives strain in the direction N. Thus the three numbers given by (3.9.13) are the extreme values of strains and since the angle between the mutually orthogonal directions N(1) directions N(1)

N(2) N(3) does not change, (3.9.13) are the principal strains and N(1) N(2) N(3) N(2) and N(3) .
APPENDIX: SOLUTION OF A CUBIC EQUATION

are the axes of principal strain. Note that there is no shear strain between any two of the three

The solution of a general cubic equation, which may have imaginary roots, in a closed form involves the use of hyperbolic functions. However, the cubic equation obtained from

det CAB ;
in which CAB

AB ] = 0

(A1)

= CBA has only real roots, and can be solved as follows.

The cubic equation

y3 + py2 + qy + s = 0
can be reduced to the form

(A2)

x3 + ax + b = 0
by substituting

(A3)

y = x + p=3:
In (A3)

(A4)

a = (3q ; p2)=3 and b = (2p3 ; 9pq + 27s)=27 :


The reduced cubic equation (A3) can be solved by transforming it to the trignometric identity

4 cos3 ; 3 cos ; cos 3 = 0:


48

(See pages 93-95, CRC Standard Mathematical Tables). The solution of the equation

3 ; I 2 + II
obtained by this method is

; III = 0

= r cos + I=3
where

cos 3 = (2I 3 ; 9(I )(II ) + 27III )=2(I 2 ; 3II )3=2

r = 2(I 2 ; 3II )1=2=3:


Example: The deformation of a body is given by

u1 = 3X12 + X2

u2 = 2X22 + X3

u3 = 4X32 + X1 :

a) Find the principal strains at the material point (1,1,1) in the reference conguration. b) Find the axis of the maximum strain through the material point (1,1,1) in the reference conguration. Also nd the direction cosines of the line into which this axis is deformed. Solution: (a) At any point

1 + 6X1 1 0 1 + 4X2 1 FiA] = 4 0 1 0 1 + 8X3


Thus

3 5

7 1 0 FiA j(1 1 1)] = 4 0 5 1 1 0 9


2

3 5 3

CAB ] = F]T F] = 4
2

7 0 1 7 1 0 1 5 0 54 0 5 1 5 0 1 9 1 0 9 3 50 7 9 7 26 5 5 : 9 5 82

32

49

The principal invariants of

C] are given by

I = CAA = 50 + 26 + 82 = 158 III = det C] = 50 26(82) ; 25] ; 7 7(82) ; 45] + 9 35 ; 234] = 99856 5 + 50 9 + 50 7 = 7377 II = det C](C ;1)AA = 26 5 82 9 82 7 26 r = 2(I 2 ; 3II )1=2 =3 = 35:484
cos 3 = (2I 3 ; 9I (II ) + 27III )=2(I 2 ; 3II )3=2 = 0:31382 3 = 71:7 360 ; 71:7 360 + 71:7 = 23:9 96:1 143:9
2 1

= 85:11

2 2

= 48:9

2 3

= 24 :

) Principal strains at the material point (1 1 1) are 42.06, 23.95, 11.5.


(b) To obtain N(1) we need to solve

(CAB ;

(1) 2 1 AB )NB

=0 (i) (ii) (iii) (iv)

(1) (1) NB NB = 1: ;35:11N1(1) + 7N2(1) + 9N3(1) = 0 7N1(1) ; 59:11N2(1) + 5N3(1) = 0 9N1(1) + 5N2(1) ; 3:11N3(1) = 0 N1(1) N1(1) + N2(1) N2(1) + N3(1) N3(1) = 1 :

7(i) + 35:11(ii) gives

; 2026:35N2(1) + 238:55N3(1) = 0
N3(1) = 8:494N2(1) :
5(i) ; 9(ii) gives

; 238:55N1(1) + 567N2(1) = 0
N1(1) = 2:377N2(1) :
50

(1) (1) Substituting for N3 and N1 into (iv), we get

(2:377)2N2(1) N2(1) + N2(1) N2(1) + (8:494)2N2(1) N2(1) = 1:

N2(1) = 0:1126
)

N3(1) = 0:957 N1(1) = 0:268 :

) The axis of the maximum strain at the material point P (1 1 1) is (0:268 0:1126 0:957).
To nd the direction cosines of the line into which this is deformed, let

PQ = ds(0:268 0:1126 0:957):


Then (P 0 Q0 )j

= FjA(PQ)A,

7 1 0 0 4 P Q ] = ds 0 5 1 0 ) Direction cosines of P0 Q0 are (0:216

0 0:268 1:989 5 4 5 4 1 0:1126 = ds 1:52 5 9 0:957 8:881 0:164 0:962).

32

Exercise: Given the displacement components

u1 = 0:1X22 u2 = u3 = 0:
a) Find the principal strains at the material point (1

1 0) in the reference conguration. 1 0) in the reference conguration.

b) Find the axis of the maximum strain at the point (1 Exercise: Given the following displacement components

u1 = 2X12 + X1 X2 u2 = X22 u3 = 0:
Find the principal strains and their axes at the material point (1=2 tion. 3.10 Deformation of Areas and Volumes Consider two different innitesimal line elements PQ and PR emanating from a point P in the reference conguation. During the deformation lines

0 0) in the reference congura-

PQ

and

PR

are deformed into

51

P0Q0 and P0R0 respectively.

PQ and PR in the reference conguration is deformed into the one with adjacent sides as P0 Q0 and P0 R0 . Let us denote the areas of these by dA and da respectively. Then
Hence the parallelogram whose adjacent sides are

dA = PQ PR
)
Also

dAB = "BCD (PQ)C (PR)D :

da = P0Q0 P0R0
dai = "ijk (P 0Q0)j (P 0R0 )k :
Recalling that (P 0 Q0 )j

= FjC (PQ)C , we obtain

dai = "ijkFjC (PQ)C FkD (PR)D


;1 ))(PQ) (PR) = "pjkFjC FkD (FpB (FBi C D

= J"BCD (F ;1)Bi(PQ)C (PR)D = J (F ;1)BidAB :


Hence

da = J (F;1)T dA:

(3.10.1)

Now consider the parallelepiped formed by three nonplanar innitesimal line elements

PR and PS passing through a point P in the reference conguration. Because of the deformation,
52

PQ,

the parallelepiped is deformed into the one whose three concurrent sides are

P0S0. If dV

P0Q0, P0R0 and

and

dv denote the volumes of these in the reference and the current congurations dV = PQ PR PS = "BCD (PQ)B (PR)C (PS )D :

respectively, then

Similarly

dv = P0Q0 P0R0 P0S0 = "ijk(P 0Q0)i (P 0R0 )j (P 0S 0)k :


Substituting

(P 0Q0 )i = FiB (PQ)B etc:

we get

dv = "ijk FiB (PQ)B FjC (PR)C FkD (PS )D


= J"BCD (PQ)B (PR)C (PS )D = J dV :
Hence

dv = J dV :
A deformation such that

(3.10.2)

dv = dV
at every material point in the body is called an isochoric (volume preserving) deformation. Thus for an isochoric deformation,

J = 1 at each material point of the body.

Examples of isochoric

deformations are the simple shearing deformation given in the example problem on page 3-2 and the one given in the exercise on page 3-31. Note that the latter is not a homogeneous deformation even though it is isochoric. Exercise: Given the following deformation

u1 = 2X12 + X1 X2
53

u2 = X22

u3 = 0:

(1=2 0 0) in the reference conguration, consider an innitesimal plane formed by the vectors 10;2 (1 0 0) and 10;2 (1 1 0). Find the (vector) area of the element into
At the material point which this plane is deformed. Exercise: The displacement components for a body are u1 At the material point (1

= 2X1 + X2 u2 = X3 u3 = X3 ; X2 .

0 0) on the surface of the body in the reference conguration, an element of area has components 10;2 (1 1 1). Find the components of the area into which this is deformed.
3.11 Mass density. Equation of continuity. Consider a sphere of innitesimal radius centered at a point P in the reference conguration. The material contained within the sphere has mass

M . Let V

denote the volume of the sphere. The

mass density 0 at point P in the reference conguration is dened as

M: = lim V !0 V M and V

(3.11.1)

One of the assumptions in continuum mechanics is that the limit on the right-hand side of (3.11.1) exists at every point of the body. Since both are positive, therefore, the mass density

0 (X1

X2 X3) is positive.

Note that from the point of view of Atomic Physics, the assumption

that the right-hand side of (3.11.1) is well dened may not be justied. One can always envisage innitesimal volume elements surrounding a point which contain no atomic particles at some time and hence will make the mass density at the point to be zero at that time. However, in continuum mechanics, we are concerned with gross effects or macroeffects of deformation, and lengths considered are much larger than the distance between adjacent atoms. We now make the assumption that the mass of the material contained in every small volume element at P is conserved. That is, the mass of the matter enclosed in the innitesimal parallelepiped at

equals the mass of the matter contained in the innitesimal parallelepiped at

P 0 into which

the former is deformed. Denoting the mass density in the present conguration by , we have

dv = 0dV:

54

Substituting for dv from (3.10.2), we arrive at

J = 0:
or

(3.11.2)

(X1 X2 X3 t)J (X1 X2 X3 t) = 0 (X1 X2 X3):


Equation (3.11.2) which relates the mass density in the present conguration to the mass density in the reference conguration is the equation of continuity or the conservation of mass in the Lagrangian description. To obtain the equation of continuity in the spatial description, we take the material derivative of (3.11.2) and thereby obtain

_J + J_ = 0:
Since

(3.11.3)

J = "ABC F1A F2B F3C


therefore,

_1A F2B F3C + F1AF _2B F3C + F1AF2B F _3C ): J_ = "ABC (F


However,

(3.11.4)

_1 F F _1A F2B F3C = "ABC @ x "ABC F @XA 2B 3C = "ABC @v1 @x1 + @v1 @x2 + @v1 @x3 F2B F3C @x1 @XA @x2 @XA @x3 @XA @v1 F F F + 0 + 0 = "ABC @x 1A 2B 3C 1 @v1 : = J @x 1
Similarly, one can show that

(3.11.5)

_2B F3C = J @v2 "ABC F1AF @x2 _3C = J @v3 : "ABC F1AF2B F @x3
55

(3.11.6)

Substituting from (3.11.5) and (3.11.6) into (3.11.4) and then the result into (3.11.3) we get

_J + J @vi = 0 @xi
and thus conclude that

@vi = 0: _ + @x i
This equation can equivalently be written as

(3.11.7)

@ + @ ( v)=0 @t @xi i
which is the continuity equation in spatial description. For an isochoric deformation

(3.11.8)

J =1
and hence

J_ = 0 _ = 0:
assumes the form

(3.11.9)

Equations (3.11.9) and (3.11.7) imply that for an isochoric deformation, the continuity equation

@vi = 0: @xi
Exercise: Show that

(3.11.10)

2x1x2 x3 v1 = ; (x 2 + x2 )2
1 2
Example: For the velocity eld given by

2 x2 1 ; x2 )x3 v2 = (( x2 + x2 )2 1 2

2 v3 = x2 x + x2 1 2

are the components of a velocity eld in an isochoric deformation.

i vi = 1 x +t t 0
nd the density of a material particle as a function of time. Solution: For the given velocity eld,

@vi = 3 : @xi 1 + t
56

Therefore, from the conservation of mass, we get

)
Integration of this equation gives

_ = ; @vi = ; 3 : @xi 1+t d = ; 3dt : 1+t ()


0 at t = 0, then

`n = ;3`n(1 + t) + A
where A is a constant of integration. If

`n 0 = ;3`n1 + A or A = `n 0 :
Thus eqn. (

) becomes = 0 =(1 + t)3:

Exercise: Given the velocity eld

vi = x1t i1 + x2t
determine how the uid density varies with time.

i2

Exercise: In the spatial description the density of an incompressible uid is given by Find a permissible form for the velocity eld, with v3 equation be satised. 3.12 Rate of Deformation Let us consider a material element P0 Q0 emanating from a material point located at P 0 (x1 in the present conguration at time t. We wish to compute length and direction of

= kx2 .

= 0, in order that the conservation of mass

D(P0Q0 ) Dt

x2 x3 )

, the rate of change of

P 0 be the deformed position of the material point located at P (X1 X2 X3) in the reference conguration and Q0 that of Q(XA + dXA). Then
Let

P0Q0.

(P 0Q0)i = xi(XA + dXA t) ; xi(XA t):

57

Thus,

= vi AdXA: D(P0Q0) in a material description. To obtain D(P0Q0) in a spatial Equation (3.12.1) expresses Dt Dt description, we note that since vi (XA t) is the velocity of a material point P presently at the position xi , therefore, if a spatial description of velocity is employed, this velocity is given by vi = vi(xj t). Note that vi (xj t) and vi(XA t) are, in general, different functions. Thus,

D (P 0Q0 ) = v (X + dX t) ; v (X t) i i A A i A Dt @vi dX ' @X A


A

(3.12.1)

D(P0Q0)i = v (x + dx t) ; v (x t) i j j i j Dt @vi dx = v dx ' @x j ij j


j

(3.12.2)

= 1=2(vi j + vj i) + 1=2(vi j ; vj i)]dxj = (Dij + Wij )dxj


(3.12.3)

where we have set

Dij = 1=2(vi j + vj i) Wij = 1=2(vi j ; vj i):

(3.12.4)

Dij , the symmetric part of the velocity gradient vi j , is known as the rate of deformation tensor or the strain-rate tensor, and Wij , the antisymmetric part of the velocity gradient vi j , is known as the
spin tensor. In the following, we give a geometric interpretation of the elements of Dij . Let

P0Q0 = dsn
where n is a unit vector in the direction of P0 Q0

= dx. Then

or

D (ds2) = D (dx dx ) = 2dx D (dx ) i Dt Dt i i Dt i D (ds) = dx (D + W )dx ds Dt i ij ij j


58

where we have substituted for

D (dx ) from (3.12.3). Since W = ;W , therefore ij ji Dt i dxiWij dxj = 0

and we get

D (ds) = dx D dx ds Dt i ij j
or

1 D(ds) = n D n : i ij j ds Dt
If n

(3.12.6) Thus D11 gives the rate of change of

= (1 0 0), the right-hand side of (3.12.6) equals D11.

length per unit length known as stretching or rate of extension for a material line presently parallel to x1 -axis. Similarly D22 and D33 give, respectively, the stretching of a material line presently in the x2 - and x3 -direction. To obtain a physical interpretation of the off-diagonal elements of let

Dij ,

P0Q0 = ds1n
Then

and

P0R0 = ds2m :

1 D (P0Q0 P0R0) = m (D + W )n + n (D + W )m i ij ij j i ij ij j ds1ds2 Dt 1 D(ds1ds2) cos (m n) ; sin (m n) _(m n) = 2mi Dij nj : ds1ds2 Dt
Here (m n) is the angle between directions m and n. Thus for m = (1

(3.12.7)

0 0) and n = (0 1 0), the right-hand side of (3.12.7) equals 2D12 and the left-hand side equals ; _(m n) . Hence 2D12 equals the rate of decrease of angle from of two line elements presently parallel to x1 and x2 -axes, 2 known as shearing or rate of shear. A similar interpretation holds for D23 and D31 . Since Dij is symmetric, we also have the result that there always exist three mutually perpendicular directions (eigenvectors of Dij ) along which the stretching (an eigenvalue of Dij ) is either
maximum or minimum among stretchings for all differential elements extending from the material particle which currently is at P 0 .
59

Note that the strain-rate tensor does not, in general, equal the time rate of change of the strain tensor. To prove this, we rst conclude from eqns. (3.12.2) and (3.7.7) that

_iA dXA = vi j FjAdXA F


which must hold for all choices of dXA . Thus

_iA(F ;1)Aj vi j = F

_ ;1 or L = FF

(3.12.8)

and the denition (3.8.6)1 of the strain tensor E gives

where L is the spatial velocity gradient. Differentiation of both sides of (3.8.2) with respect to time

_ = 2C _ =F _ T F + FT F _ = FT (LT + L)F E = 2FT DF :


the strain tensor E.

(3.12.9)

_ 6= D, and the strain-rate tensor D should not be confused with the time rate of change of Hence E
We now attempt to provide a physical interpretation of the spin tensor. Let n be a unit eigen-

vector of have

D, i.e., Dn = n where

is the eigenvalue corresponding to n. From (3.12.3), we

_ i + dsn (dsni) = dsn _ i = (Dij + Wij )dsnj

n _ i = Dij nj + Wij nj ; (nk Dkl nl )ni


= Wij nj
(3.12.10)

where we have used (3.12.6). It states that the spin tensor operating on a unit eigenvector of

gives the rate of change of that unit vector. Thus the spin equals the angular velocity of the principal axes of stretching. The axial vector

wi = ; ijk Wjk

or

w = curlv

is the vorticity vector; its direction is the axis of spin, and its magnitude is the vorticity magnitude,
60

w. w = pwiwi = sWjk Wjk :


p

We now show that the spin tensor dened by (3.12.4)2 does not equal the rate of change of the rotation matrix R. Equations (3.12.9) and (3.13.1)1 give

_ + RU _ )(U;1 RT ) L = (RU
or

(3.13.15)

_ T + RUU _ ;1 RT D + W = RR _ ;1 + U;1 U _ )RT + 1 R(UU _ ;1 ; U;1 U _ )RT _ T + 1 R(UU = RR 2 2


where we have added and subtracted skew-symmetric tensors on both sides, we obtain

(3.13.16)

1 RU;1UR _ T to the right-hand side. Equating symmetric and 2

_ and W 6= R _. These equations clearly evince that D 6= U

;1 _ T _ ;1 D=1 2 R(UU + U U)R _ ;1 ; U;1U _ )RT : _ T + 1 R(UU W = RR 2

(3.13.17)

Taking the material derivative of both sides of eqn. (3.10.2) we obtain

D (dv) = J_ dV Dt @vi dV = J @x i = @vi dv @xi


and, therefore,

(3.12.11) (3.12.12)

1 = @vi = D = I : ii D dv dv @xi
In going from (3.12.11) to (3.12.12) we used

(3.12.13)

@vi = JI J_ = J @x D
i
61

which can be obtained from eqns. (3.11.4), (3.11.5) and (3.11.6). Thus the rst principal invariant

ID of the rate of deformation tensor Dij gives the rate of change of volume per unit volume.
We now derive an expression for the rate of change of an area element. Rewriting eqn. (3.10.1) as FT da = JdA and taking the time derivative of both sides, we obtain

_ T da + FT d_a = Jd _A F FT LT da + FT d_a = JID dA _ i = ID ij ; @vj daj : d_a = (ID ; LT )da da @xi _ = 0 but d_a 6= 0 in general. In an isochoric deformation, dv
Example: Given the velocity eld or (3.12.14)

vi = 2x2
nd (a) the rate of deformation and the spin tensors,

1i

(b) the rate of extension per unit length of the line element P0 Q0 (c) the maximum and the minimum rates of extension. Solution (a) The matrix of the velocity gradient is

= 10;2(1 2 0 ),

0 2 0 vi j ] = 4 0 0 0 0 0 0
Therefore,
2

3 5

:
3

0 1 0 Dij = 1=2 vi j + vj i] = 4 1 0 0 5 0 0 0 2 3 0 1 0 Wij = 1=2 vi j ; vj i] = 4 ;1 0 0 5 : 0 0 0


62

(b) Given P0Q0

p 1 2 p 0 = 10;2(1 2 0) = 10;2( 5) p 5 5
1 D(ds) = n D n = fngT D]fng i ij j ds Dt 2 p p 40 1 0 = (1= 5 2= 5 0) 1 0 0 0 0 0

. Thus n = (1=

5 2= 5 0). From

eqn. (3.12.6)

32 54

1=p5 2= 5 5 = 4=5 : 0

(d) From the characteristic equation

det Dij ;

ij ] = 0

1. Therefore, 1 is the maximum and ;1 the minimum rate of extension. The eigenvector n(1) (for 1 = 1) determined from

we determine the eigenvalues of the tensor Dij as

=0

Dij n(1) j ;
and is n(1)

(1) 1 ij nj

=0 = ;1 is (1= 2)(1 ;1 0). The

= (1= 2)(1 1 0). Similarly n(2) corresponding to third eigenvector is n(3) = (0 0 1).
Exercise: For the velocity eld

(1) n(1) j nj = 1 2

vi = 2x2 2
nd

i1

a) the rate of extension per unit length of a material line element which in the present conguration is given by 10;2 (1

1 0) through the point (5 3 0), = Xi at time t = 0,

b) the deformation corresponding to the given velocity eld if xi

c) the principal stretches and the deformed position of the axis of the maximum principal stretch for the material particle which at t = 1 is at (0

1=2 0), 1=2 0).

d) principal stretchings and their axes for the material particle which currently is at (0
63

3.13 Polar Decomposition The polar decomposition theorem of Cauchy3 states that a non-singular matrix equals an orthogonal matrix either pre or post multiplied by a positive denite symmetric matrix. If we apply this theorem to the deformation gradient F, we get

F = RU = VR

(3.13.1)

Note that the decomposition (3.13.1) of F is unique in that R, U and V are uniquely determined by F. From (3.13.1) it follows that

in which R is a proper orthogonal matrix and U and V are positive denite symmetric matrices.

J = det F = det U = det V:


Since C = FT F and B = FFT , therefore

(3.13.2)

C = FT F = (RU)T RU = U2 B = FFT = VR(VR)T = V2:


We note that

(3.13.3) (3.13.4)

V = RURT
and

(3.13.5)

B = FFT = RU(RU)T = RCRT : U is symmetric, it has at least one orthogonal triad of eigenvectors. eigenvector of U and 1 be the corresponding eigenvalue so that
Since Let

(3.13.6)

N(1) be an
(3.13.7)

UN(1) = 1N(1) :
Therefore

CN(1) = UUN(1) = 1UN(1) = ( 1 )2N(1) :


64

(3.13.8)

3 This theorem is proved in any book on linear algebra, e.g. on pg. 83 of P.R. Halmos, Finite Dimensional Vector Spaces, 2nd ed. Van Nostrand, Princeton, Toronto, and London, 1958.

holds for every eigenvector of U, we conclude that eigenvectors of U and C are the same and the eigenvalues of C are equal to the squares of the eigenvalues of U. In section 3.9, we proved that the eigenvalues of C are the squares of the principal stretches and the eigenvectors of C are the axes of principal stretches also usually called principal axes of stretch in the reference conguration. Thus eigenvectors of eigenvalues of U are the princpal stretches. Let us now nd the deformed position of an eigenvector

Thus N(1) is an eigenvector of C and the corresponding eigenvalue is

( 1)2.

Since eqn. (3.13.8)

U are the principal axes of stretch in the reference conguration and the

N(1) of U. Since

FN(1) = RUN(1) = 1RN(1)


therefore

(3.13.9)

n(1) = RN(1)
points in the direction of the vector into which N(1) is deformed. Note that

(3.13.10)

n(1) n(1) = n(1)T n(1) = N(1)T RT RN(1) = N(1)T N(1) = 1


therefore, n(1) is a unit vector in the direction of the deformed position of N(1) . Now

(3.13.11)

FN(1) = VRN(1) = Vn(1) :


Equations (3.13.9), (3.13.10) and (3.13.12) when combined together give

(3.13.12)

Vn(1) = 1 n(1) :

(3.13.13)

of U and V are equal and that the eigenvectors of V are the deformed positions of the eigenvectors of

Thus 1 is an eigenvalue of V with n(1) as the eigenvector. This exercise proves that the eigenvalues

V are the deformed position of the principal axes of stretch. Said differently, eigenvectors of V are the principal axes of stretch in the deformed conguration or in
Thus eigenvectors of

U.

the present conguration. Of course,

Bn(1) = VVn(1) = ( 1)2 n(1) :


65

(3.13.14)

From (3.13.14), (3.13.8), (3.13.9) and (3.13.10), we conclude that eigenvalues of

equal, and the eigenvectors of B are the principal axes of stretch in the deformed conguration. of the triad N(1)

B and C are

Corresponding to the two decompositions of F given by (3.13.1) we can view the deformation

N(2) N(3) , the eigenvectors of U, as a stretch of these axes followed by a rotation

or a rotation of these axes followed by their stretch. This is schematically shown in the Fig. below.

Example: For simple shear

xi = Xi + kX1
a) nd the principal stretches,

2i

b) show that the angle through which the principal axes of stretch in the reference conguration are rotated so as to become the principal axes of stretch in the present conguration is given by tan

= k=2.

66

Solution: For the given deformation


2

F] = 4
2

C] = 4
2

B] = 4
The squares of the principal stretches 1

1 0 0 k 1 05 0 0 1 1 + k2 k k 1 0 0 1 k k 1 + k2 0 0

0 0 1 0 0 1

3 5 3 5

3 are roots of the equation det( C] ;

1]) = 0

(1 + k2 ; )(1 ; ) ; k2 ](1 ; ) = 0:
p

) ( 1 )2 = 1 + 1=2k2 + k 1 + 1=4k2

( 2)2 = 1 + 1=2k2 ; k 1 + 1=4k2 = 1=( 1)2


3

= 1:

Note that the given deformation is a plane strain deformation. By looking at the matrices and eigenvalue one. Also, x3 -axis coincides with the

B] we see that X3-axis and x3 -axis are eigenvectors of C] and B] corresponding to the

C]

X3-axis for the given deformation. Thus the rotation of principal axes of stretch takes place about the X3 -axis. Let this angle of rotation be in
the clockwise direction. Then

cos 4 R] = sin 0
Substituting for

; sin 0
cos 0 0 1

3 5

B], C] and R] in eqn. (3.13.6) we arrive at 3 2 3 2 0 1 k 0 1 + k2 cos2 ; 2k sin cos 4 k 1 + k 2 0 5 = 4 k (cos2 ; sin2 ) + k 2 sin cos 1 + k2 sin2 + 2k sin cos 0 5
0 0 1 0 0 1

Therefore

1 = 1 + k2 cos2 ; 2k sin cos


which gives tan

= k=2.
67

3.14 Innitesimal Deformations When the displacements and displacement gradients are small, we can neglect second order terms in (3.8.6) and approximate the strain tensor by

eAB = 1=2(uA B + uB A):


Since

(3.14.1)

@uA = @uA @xj @XB @xj @XB @uj A = @u jB + @xj @XB @uA @uj A = @u jB + @xj @xj @XB A ' @u @xj jB
deformations,

(3.14.2)

where we have neglected the second order term in the displacement gradients; for innitesimal

@uA + @uB = 1 @ui + @uj eAB = 1 2 @XB @XA 2 @xj @xi


coordinates to evaluate the innitesimal strain tensor.

iA jB :

(3.14.3)

Thus one can differentiate displacements with respect to the current coordinates or the referential Since eAB is symmetric, therefore, it has at least one orthogonal triad of eigenvectors. The eigenvectors of eAB are the principal axes of engineering strain and its eigenvalues are principal innitesimal strains. From equations (3.8.3) and (3.8.6) we obtain
p p jP0Q0j=jPQj = 1 + 2E11 ' 1 + 2e11 ' 1 + e11 :

(3.14.4)

= ds (1 0 0) we see that e11 equals the change in length per unit length of an innitesimal line element parallel to X1 -axis. Similar interpretation holds for e22 and e33 . From
equations (3.8.4) and (3.8.6) we conclude that

Recalling that PQ

cos = p

2Ep 2e12 12 ' ' 2e12 : 1 + 2E11 1 + 2E22 (1 + e11 )(1 + e22 )
68

(3.14.5)

Here is the angle between the deformed positions of lines initially parallel to X1 and X2 axes. If the change in the angle is small so that

=2;

, then eqn. (3.14.5) becomes (3.14.6)

sin ' = 2e12 :


and X2 axes. This change in the angle is called the shearing strain. From equations (3.9.3) and (3.8.6) we obtain

Thus 2e12 equals the innitesimal change in the angle between two lines originally parallel to X1

jP0Q0j = 1 + e N N : (3.14.7) AB A B jPQj Thus the engineering strain in any direction N equals NA eAB NB . One can similarly show that the shearing strain between two orthogonal directions M and N equals 2MA eAB NB . Recalling that
FiA =
and that

iA + ui A

iA + HiA

CAB = FiAFiB
we obtain

C] = F]T F] ' 1] + H] + H]T 1 ( H] + H]T ) U] = C]1=2 ' 1] + 2 = 1] + e] R] = F] U];1 = ( 1] + H])( 1] + e]);1 1 ( H] ; H]T ) : ' 1] + 2
The skew symmetric tensor

(3.14.8)

(3.14.9)

(3.14.10)

gives the innitesimal rotation. For any innitesimal vector PQ we have

1 (u ; u ) !AB = 2 AB BA

(3.14.11)

fP0Q0g = F]fPQg = ( 1] + e] + !])fPQg


69

and therefore

fP0Q0g ; fPQg = e]fPQg + !]fPQg :


Thus the deformation of a line element innitesimal strain tensor and the innitesimal rotation tensor. Similarly, if

(3.14.12)

PQ equals the sum of the deformations caused by the u = u1 + u2


(3.14.13)

then

e] = e1 ] + e2 ]
!] = ! 1 ] + ! 2 ] :

(3.14.14) (3.14.15)

Thus the innitesimal strains and rotations caused by a given displacement equal the sum of the innitesimal strains and rotations caused by the components of the given displacement. Exercise: Prove that for small deformations

J = 1 + eAA
= 0(1 ; eAA)
and for isochoric deformations

(3.14.16) (3.14.17)

eAA = 0:
Exercise: Consider the displacement eld

(3.14.18)

uA = k (2X12 + X1X2 )
unit length for the material line element PQ = ds (1

2 A1 + X2 A2 ]:

Using both the innitesimal strain theory and the nite strain theory, nd the change in length per

1 0) that emanates from the material particle P (1 1 1) in the reference conguration for k = 10;4 10;3 10;2 10;1 1 10. Plot these changes as a function of k.
Exercise: With reference to a rectangular Cartesian coordinate system, the state of strain at a point

70

is given by the matrix.

a)

5 3 0 ; 4 4 e] = 10 3 4 ;1 5 : 0 ;1 2 What is the engineering strain in the direction 2e1 + 2e2 + e3 at the point (1 1 1) in the
reference conguration?

b) What is the shearing strain between two perpendicular lines (in the reference conguration) emanating from the point (1

1 1) in the directions of 2e1 + 2e2 + e3 and ;3e1 + 6e3?

For an innitesimal rigid body motion, the displacement vector u is given by

uA = cA + bAB XB
in which cA is a constant and bAB by (3.14.19), eAB

(3.14.19)

= ;bBA is a skew-symmetric tensor. For the displacement given

= 0. Naturally the following question arises: Is the rigid body motion the only

deformation for which the innitesimal strain tensor vanishes identically? The answer, as proved in the following example, is yes. Example: Regarding eAB Solution: eAB

= 0 as six partial differential equations, solve for u.

= 0 corresponds to

@u1 = @u2 = @u3 = 0 @X1 @X2 @X3


and

(3.14.20)

@u1 + @u2 = @u2 + @u3 = @u3 + @u1 = 0: (3.14.21) @X2 @X1 @X3 @X2 @X1 @X3 By differentiating (3.14.21)1 with respect to X2 and (3.14.21)3 with respect to X3 and making use
of (3.14.20)2 3 we obtain

@ 2 u1 = @ 2 u1 = 0 : @X22 @X32
This when combined with (3.14.20)1 yields

u1 = c1 + b12 X2 + b13 X3 + a1 X2X3


71

(3.14.22)

in which c1

b12 b13 and a1 are constants. By following a procedure similar to that used to obtain (3.14.22) for u1 , we obtain for u2 and u3 the following. u2 = c2 + b21 X1 + b23 X3 + a2 X1X3 u3 = c3 + b31 X1 + b32 X2 + a3 X1X2:
Substituting from (3.14.22), (3.14.23) and (3.14.24) into (3.14.21) we get (3.14.23) (3.14.24)

b12 + b21 + (a1 + a2 )X3 = 0 b23 + b32 + (a2 + a3 )X1 = 0 b31 + b13 + (a3 + a1 )X2 = 0:
Since these equations hold for all values of X1 body, therefore,

X2 X3 which correspond to various points in the

b12 + b21 = b23 + b32 = b31 + b13 = 0 a1 + a2 = a2 + a3 = a3 + a1 = 0:


The last set of equations implies that a1 that bAB

= a2 = a3 = 0. From the other set of equations, it follows

= ;bBA .

Hence equations (3.14.22), (3.14.23) and (3.14.24) which are a solution of

eAB = 0 reduce to uA = cA + bAB XB :


and u1 + u2 is the same if u2 is a rigid body motion. To prevent the rigid motion of a body, one needs to x three noncolinear points of the body. The strain eld given by eqns. (3.14.20) and (3.14.21) is a very special one in that it is identically zero throughout the body. What if we were given Recalling equations (3.14.13) and (3.14.14) we see that the strain caused by displacements u1

e11 = f (X1 X2) e22 = g(X1 X2) e12 = h(X1 X2)


72

(3.14.25)

and asked to nd the corresponding two-dimensional displacement eld? Can we always nd a displacement eld that will produce the strains specied by (3.14.25)? The answer is of course no. Since

@u1 e = @u2 2e = @u1 + @u2 e11 = @X 22 @X2 12 @X2 @X1 1


therefore,

@ 2 e12 = @ 3 u1 + @ 3 u2 = @ 2 e11 + @ 2 e22 : 2 @X @X22 @X1 @X12 @X2 @X22 @X12 1 @X2
If we substitute for e11

e22 and e12 from (3.14.25) we obtain @2h = @2f + @2g : 2 @X @X22 @X12 1 @X2
(3.14.26)

Thus unless the given functions f

g, and h satisfy (3.14.26) we will not be able to nd a displace-

ment eld that will produce the desired strain eld.4 Another way of saying essentially the same thing is that we have three equations (3.14.25)1 2 3 for two unknowns u1 and u2 . Unless the given expressions for e11 , e22 and e12 are related somehow, we will not, in general, be able to nd u1 and u2 . That relation is the equation (3.14.26) which is known as a compatability condition. In the three dimensional case we can derive compatability conditions like (3.14.26) in a similar way. However, another, perhaps neater, approach to the problem is the following. of P , we would like to nd the displacement uQ of a neighboring point Q. Now Given the displacement uP of a point P in the body and the strain eld eAB in the neighborhood
Z Q Z Q

P uQ A ; uA

= =

P Z Q P

duA =

uA B dXB
(3.14.27)

(eAB + !AB )dXB :

Note that
Z Q

!AB dXB =

Z Q

Q) !AB d(XB ; XB Q P Q )! ; (XB ; XB AB C dXC P


Z

Q) = !AB (XB ; XB
4 This

(3.14.28)

requirement is in addition to the condition that J

> 0 for the given strain eld.


73

where we have integrated by parts. Combining (3.14.27) and (3.14.28) we obtain Z Q Q Q P P P uA = uA ; !AB (XB ; XB ) + RAC dXC P where

(3.14.29)

Q )! RAC = eAC ; (XB ; XB AB C :

(3.14.30)

Since the path of integration in (3.14.29) from point P to point Q is arbitrary, in order to obtain a unique value of uQ , the integral in eqn. (3.14.29) must be path independent. The necessary and sufcient condition for this is

RAC D = RAD C :
Substituting from (3.14.30) into (3.14.31) and noting that

(3.14.31)

1 !AB C = 1 ( u A BC ; uB AC ) + (uC AB ; uC AB ) = eAC B ; eBC A 2 2


we arrive at

eAB CD + eCD AB ; eAD BC ; eBC AD = 0: e11 22 + e22 11 ; 2e12 12 = 0 e22 33 + e33 22 ; 2e23 23 = 0 e33 11 + e11 33 ; 2e13 13 = 0 e11 23 + e23 11 ; e13 12 ; e12 13 = 0 e22 31 + e31 22 ; e21 23 ; e23 21 = 0 e33 12 + e12 33 ; e32 31 ; e31 32 = 0:

(3.14.32)

Even though there are 81 equations given by (3.14.32) only the following six are non-trivial.

(3.14.33)

Out of these six equations only three are linearly independent. However, we will not prove that here. Example: For the two-dimensional small strain theory, the strains for a cantilever beam are given by

e11 = AX1X2 e22 = ; AX1 X2 2e12 = A(1 + )(a2 ; X22 )


74

(3.14.31)

where A, a, axes X1

are positive constants and

X2 are functions of X1 X2.

1 . Assume that the displacements u , u relative to 1 2 2

a) Show that continuous single-valued displacements (u1 b) Hence derive formulas for (u1

u2) are possible.

u2) as explicit functions of (X1 X2) with the conditions u1 = u2 = u1 2 = 0 for X1 = L X2 = 0.

Solution: a) From the given expressions for e11

e22 and e12 we obtain J = 1 + e11 + e22 = X1 and X2, J > 0. Thus continuous

e11 22 = e22 11 = e12 12 = 0


so that the only non-trivial compatability equation (3.14.33)1 is satised.

1 + A(1 ; )X1 X2.


b) Integrating

For the given constraints on A

single-valued displacements are possible.

@u1 = AX X @u2 = ; AX X 1 2 1 2 @X1 @X2


we obtain

u1 = A X 2X + f (X2) u2 = ; 2 AX1 X22 + g(X1): 2 1 2


Substituting for u1 and u2 into

@u1 + @u2 = A(1 + )(a2 ; X 2 ) 2 @X2 @X1


we obtain

df ; AX 2 + A(1 + )X 2 = A(1 + )a2 ; dg ; A X 2: 2 dX2 2 2 dX1 2 1


75

Since the left-hand side of this equation is a function of X2 and the right-hand side a function of

X1, for the two sides to be always equal, each must equal a constant say c. Thus df ; AX 2 + A(1 + )X 2 = c 2 dX2 2 2 dg ; A X 2 = c: A(1 + )a2 ; dX 2 1
1
Therefore

f = 6 AX23 ; A (1 + )X23 + cX2 + d 3 g = A(1 + )a2 X1 ; A X 3 ; cX1 + e 6 1


where d and e are constants of integration. Thus

2 X + AX 3 ; A (1 + )X 3 + cX + d u1 = A X 2 2 1 2 2 6 2 3 3 u2 = ; 2 AX1X22 + A(1 + )a2 X1 ; A 6 X1 ; cX1 + e: In order that u1 = u2 = u1 2 = 0 for X1 = L X2 = 0 we must have 2 e = ;A(1 + )a2 L + A L3 ; A L3 : L d = 0 c = ;A 2 6 2

(3.14.32)

Substituting for d

For points on the plane X2

c and e into (3.14.32) we arrive at the following. 2X ; A 1 + 3 ; A L2 X u1 = A X X 2 2 1 2 2 3 2 2 2 3 ; X 3 ) + PAL (X ; L): u2 = ; 2 AX1X22 + A(1 + )a2 (X1 ; L) + A ( L 1 1 6 2
= 0,
3 ; X 3 ) + AL (X ; L) : u1 = 0 u2 = A(1 + )a2 (X1 ; L) + A ( L 1 6 2 1 2

(3.14.34)

(3.14.33)

Thus longitudinal lines on the plane X2

= 0 which is the neutral surface are not stretched. How-

ever, points on the longitudinal line do move vertically. Equation (3.14.34) corresponds to the deection equation usually studied in the rst Strength of Materials or Mechanics of Deforms course. However, in that course the following boundary conditions are used. For

X1 = L X2 = 0 u1 = u2 = u2 1 = 0:
76

With these, eqn. (3.14.32) gives

2 A A AL 2 2 3 u1 = 2 X1 X2 + A(1 + )a X2 ; 3 1 + 2 X2 ; 2 X2 2 3 3 + AL X ; AL : X u2 = ; 2 AX1 X22 ; A 1 6 1 2 3


Thus

u2(0 0) = ; AL 3
which agrees with the strength of materials solution.

Exercise: Check whether or not the following distribution of the state of innitesimal strain satises the compatability conditions:

X12 X22 + X32 X1X3 2 0 X1 5 : e] = 4 X2 + X32 X1 X3 X1 X22


Exercise: Given the strain eld

e12 = e21 = X1X2


and all other eAB

= 0,

a) Does it satisfy the equations of compatability? b) By attempting to integrate the strain eld, show that it cannot correspond to a displacement eld.

4 The Stress Tensor


4.1 Kinetics of a Continuous Media In this section we will study the laws of motion applicable to a continuous medium similar to Newtons laws of motion studied in particle mechanics. We rst review Newtons laws of motion below. 1. Newtons rst law of motion: a free particle continues in its state of rest or of uniform motion.
77

2. Newtons second law of motion: In an inertial frame, the rate of change of linear momentum of a particle equals the resultant force acting on the particle. That is

d (mv) = ma: F = dt

(4.1.1)

3. Newtons third law of motion: To every action there is an equal and opposite reaction. Newtons rst law of motion denes an inertial frame. That is, an inertial frame is one in which Newtons rst law of motion holds. Usually, it is taken as a frame attached to the Sun. However, in most engineering problems, one can take the coordinate axes xed to the earth as an inertial frame without introducing any appreciable error. Hereafter, we will take an inertial frame as the frame of reference. To write the laws of motion for a continuum we note that the linear momentum of the material

dv)v where is the mass density and v is the velocity. R Hence the linear momentum of the shaded portion is ( dv )v in which the integration is over the
enclosed in an innitesimal volume dv is ( shaded region. To nd the resultant force acting on this region of interest, we observe that we have two kinds of forces.

78

Body forces are forces that act on all particles in a body as a result of some external body or effect not in direct contact with the body under consideration. An example of this is the gravitational force exerted on a body. This type of force is dened as a force intensity per unit mass or per unit volume at a point in the continuum. Thus, if the body force per unit mass is gi , then the body force on the material enclosed in the shaded region will be dm gi . Surface forces are contact forces that act across a surface of the body, which may be internal or external. In non-polar continuum mechanics we assume that the action of that part of the body which is exterior to the shaded region on the body enclosed in the shaded region is equipollent to a system of forces acting on the bounding surface of the shaded region. The assumption that the contact force is of this kind is the cut principle of Cauchy: Within the shape of a body at any given time, conceive a smooth, closed diaphragm; then the action of the part of the body outside that diaphragm and adjacent to it on that inside is equipollent to that of a eld of vectors dened on the diaphragm. Note that no point moments are assumed to be exerted by one part of the body on its adjacent part across the common surface. Thus in nonpolar continuum mechanics, moments are caused by the forces. The contact force at a point P on a surface is usually given as a force f acting on a unit area surrounding P and lying on the surface. Through a given point in the body, there are innitely many surfaces. The intensity of the contact force at the point P on each of these surfaces will, in general, be different. How does f at the point P depend upon the surface through

P?

In the classical continuum mechanics, it is assumed that the intensity of the contact force on

all surfaces with a common tangent plane at P is the same. That is, f at P is assumed to depend upon the surface through P only through the oriented normal n of the surface at P .

f = f (xi nj ):
positive. Thus f (x

(4.1.2)

This is Cauchys Postulate. A unit normal to the surface which points out of the body is taken as

n) is the intensity of the contact force at P which the unshaded portion of the body exerts on the shaded portion and f (x ;n) is the intesntiy of the contact force at P which
the shaded portion exerts on the unshaded one. The intensity of the contact force is also known as surface traction or traction or stress vector. Denoting the magnitude of the element of area on
79

the surface by da, the total contact force on the shaded region is material contained in the shaded region takes the form

f1 da. Thus eqn. (4.1.1) for the

d Z dm v = Z f da + Z dm g : i i i dt
Equation (4.1.3) is known as the conservation of linear momentum. A similar equation

(4.1.3)

d Z " x (dm v ) = Z " x f da + Z " x (dm g ) k ijk j k ijk j k dt ijk j

(4.1.4)

for the moment of momentum is known as the conservation of moment of momentum. Equations (4.1.3) and (4.1.4) are the BASIC LAWS OF MOTION of Continuum Mechanics, as far as this course is concerned.

d (dm) = 0, therefore, equations (4.1.3) and (4.1.4) can also be written as Since dt
Z Z

dm ai =

fi da + dm gi
Z

(4.1.5)
Z

"ijk xj (dm ak ) =

"ijk xj fk da + "ijkxj (dm gk ):

(4.1.6)

height " with the top and bottom faces perpendicular to n and the point P lying on one of the end faces. We apply the balance of linear momentum (4.1.5) to the

We now study the dependence of f upon n in some detail. Consider a cylinder of radius R and

material

contained within this cylinder. Using the mean-value theorem of calculus, we obtain

R2 "

)ai =

fi(x n)da +

fi(x ;n)da +

fi(x n)da + ( R2" )gi :

(4.1.7)

Top Surface

Bottom Surface

mantle

80

Here ai and gi denote the values of ai and gi evaluated at some point in the cylinder. Let the height

" of the cylinder go to zero and assume that the elds ai and gi are bounded. Then in the limit, eqn.
(4.1.7) becomes

0=

fi (x n) + fi(x ;n)]da:

Since this equation has to hold for all values of R, therefore, the integrand must be zero. That is

fi(x ;n) = ;fi (x n):

(4.1.8)

orthogonal, the fourth having outward unit normal n.

now show that f is in fact linear in n. Consider a tetrahedron, three sides of which are mutually

This is known as Cauchys Fundamental Lemma and states that f is an odd function of n. We

Let the area of the inclined plane are

ABC be A.

Then the areas of planes

PBC , PAC and PAB

An1 , An2 and An3 respectively.

On applying eqn. (4.1.5) to the material contained in the

tetrahedron, and using the mean-value theorem, we obtain

V ai = fi(x n)A + fi (x ;e1)An1 + fi (x ;e2)An2 + fi (x ;e3)An3 + V gi :

(4.1.9)

In eqn. (4.1.9), the superimposed bars indicate quantities evaluated at some point in the tetrahedron or at some point on a plane bounding the tetrahedron. In eqn. (4.1.9), dividing throughout by A, taking the limit as the tetrahedron shrinks to the point P , and using (4.1.8) we arrive at

fi (x n) = fi(x e1 )n1 + fi (x e2)n2 + fi(x e3)n3


81

(4.1.10)

where all stress vectors are evaluated at the point P . Setting n = vectors, we obtain

p + q where p and q are unit

fi(x p + q) = fi(x e1)p1 + fi(x e2)p2 + fi(x e3)p3 ]


+ fi(x e1)q1 + fi(x e2 )q2 + fi (x e3)q3 ] = fi(x p) + fi(x q) :
Thus f is a linear function of n and we can write (4.1.11)

fi (x n) = Tij (x)nj :
By comparing the right-hand sides of (4.1.10) and (4.1.12) we get

(4.1.12)

Tij (x) = fi(x ej ):


Thus Ti1

(4.1.13)

Ti2 and T13 denote, respectively, the surface tractions on planes whose outer normal points in the positive x1 x2 and x3 -directions. Tij is called the stress tensor. The plane whose outer normal points in the positive x1 -direction is simply known as the x1 -plane. Thus Ti1 Ti2 , and Ti3 are, respectively, the surface tractions on the x1 x2 and x3 -planes. Since T11 (x) = f1 (x e1) T21 = f2(x e1 ) T31 = f3 (x e1) T11 T21 and T31 are, respectively, the normal and shearing stresses on the x1 -plane. Note that the p 2 + T 2 . Also the positive values of T11 T21 and T31 resultant shear stress on the x1 -plane is T21 31 point in the positive direction of the axes on the positive x1 -plane. Because of (4.1.8), positive values of T11 T21 and T31 point in the negative direction of the axes on the negative x1 -plane.
From equation (4.1.13) it is clear that stress vectors on three mutually perpendicular planes at a point determine the stress tensor at that point. Because of eqn. (4.1.12) or (4.1.10), stress vectors on three mutually perpendicular planes at a point also determine the stress vector on any other plane. This proves Cauchys fundamental theorem: From the stress vectors acting on three mutually perpendicular planes at a point, stress vectors on every plane through the point can be determined; they are given by (4.1.12) as linear functions of the stress tensor Tij .
82

A glance at equations (4.1.5) and (4.1.6) reveals that one integration in each equation is over the surface area whereas others are over the region under consideration. We now transform this surface integral into the volume integral by using the divergence theorem. Note that
Z Z Z

fida =
Z Z

Tij nj da =

Tij j dv
Z

and
Z

"ijk xj fk da =
=
Z

"ijk xj Tkpnpda = ("ijk xj Tkp) p dv "ijk ( jpTkp + xj Tkp p)dv:

Thus equations (4.1.5) and (4.1.6) can be written as

( ai ; Tij j ; gi)dv = 0

and
Z

"ijk xj ( ak ; Tkp p ; gk ) ; Tkj ]dv = 0:

Since both these equations must hold for every region in the body, therefore, if the integrand is continuous throughout the body, then it must vanish. This gives

ai = Tij j + gi "ijk Tkj = 0 or Tij = Tji:

(4.1.14) (4.1.15)

These are Cauchys laws of motion. Equation (4.1.14) expresses the balance of linear momentum and eqn. (4.1.15) the balance of moment of momentum. We remark that these hold in an inertial frame. Thus on the assumption that the balance of linear momentum is satised, the balance of moment of momentum reduces to the requirement that the stress tensor be symmetric. In classical continuum mechanics, Tij is always taken to be symmetric so that the balance of moment of momentum is identically satised. For static problems ai

= 0 and eqn. (4.1.14) gives

Tij j + gi = 0
83

(4.1.16)

as the three equations of equilibrium. If a given stress eld satises eqn. (4.1.16), we may or may not be able to produce that stress eld statically in a continuous body. However, if a given stress eld does not satisfy even only one of the three equations of equilibrium, then it certainly cannot be produced statically in a continuous body. Example: Show that the following stress eld

2 2 T11 = x2 2 + x1 ; x2 T12 = ;2 x1 x2 2 2 T22 = x2 1 + x2 ; x1 T23 = T13 = 0


; ; 2 x + x2

T33 =
Solution

satises equations of equilibrium with zero body forces.

T11 1 = 2 x1 T12 2 = ;2 x1 T13 3 = 0 T21 1 = ;2 x2 T22 2 = 2 x2 T23 3 = 0 T31 1 = 0 T32 2 = 0 T33 3 = 0:


Thus

T11 1 + T12 2 + T13 3 = 2 x1 ; 2 x1 + 0 = 0 T21 1 + T22 2 + T23 3 = ;2 x2 + 2 x2 + 0 = 0 T31 1 + T32 2 + T33 3 = 0 + 0 + 0 = 0


and the equations of equilibrium

Tij j = 0
with zero body force are satised. Exercise: Suppose that the body force is g = ;g e3 , where g is a constant. Consider the following stress tensor

x2 ;x3 0 T] = 4 ;x3 0 ;x2 0 ;x2 T33


84

3 5

Find an expression for T33 so that Tij satises the equations of equilibrium. Exercise: Suppose that the stress distribution has the form (called plane stress)

T11 (x1 x2 ) T12 (x1 x2) 0 Tij ] = 4 T12 (x1 x2 ) T22 (x1 x2) 0 5 : 0 0 0
(a) What are the equilibrium equations in this special case? (b) If we introduce a function

(x1 x2) such that


22

T11 =

T22 =

11

T12 = ;

12

will this stress distribution be in equilibrium with zero body force? 4.2 Boundary Conditions for the Stress Tensor Applied distributed forces on the surface of a body are called surface tractions. We wish to nd the relation between the surface tractions and the stress eld dened within the body. This relation is eqn. (4.1.12) in which the left-hand side represents the surface tractions applied to the bounding surface of the body. The equation

Tij (x)nj = fi(x)


for points x on the bounding surface of the body is called the boundary condition. Example: A long prismatic dam is subjected to water pressure that increases linearly with the depth. The dam has thickness

2b and height h.

Write the traction boundary conditions for the

traction-type bounding surfaces of the dam.

85

Solution: At the top surface therefore

n = (;1 0 0).

Since there is no applied force on this surface,

Tij nj = ;Ti1 = 0
On the plane x2

on the plane

x1 = 0:

= ;b n = (0 ;1 0). There is no force applied on this surface, therefore,

Tij nj = ;Ti2 = 0
On the plane x2

on the plane

x2 = ;b:

= b n = (0 1 0). The water pressure at any point on this plane exerts a normal force equal to ; w gx1 e2 in which w is the mass density of water, g is the gravitational constant.
Therefore,

Tij nj = Ti2 = ; w gx1

i2 on the plane

x2 = b:

The lower surface of the dam is in contact with the ground and the boundary conditions on it depend upon whether the ground is taken as deformable or rigid; we will not discuss these boundary conditions here. Example: Write the traction boundary conditions at the inner and the outer surfaces of a cylindrical pressure vessel subjected to an internal pressure p1 and external pressure p2 .

Solution: At a point on the inner surface n =

; x 1

; a ; xa 0 . Therefore,
2

Tij nj = fi = ;p1 ni
86

simplies to

T11 x1 + T12 x2 = ;p1 x1 T21 x1 + T22 x2 = ;p1 x2 T31 x1 + T32 x2 = 0:


At a point on the outer surface n outer surface are

;x

x2 b

. Therefore, the traction boundary condition on the

T11 x1 + T12 x2 = ;p2 x1 T21 x1 + T22 x2 = ;p2 x2 T31 x1 + T32 x2 = 0:


Exercise: Given the following stress distribution

nd T12 in order that the stress distribution is in equilibrium with zero body force, and that the stress vector on the plane x1 = 1 is given by (1 + x2 )e1 + (5 ; x2 )e2 ]103 . Exercise: Consider the following stress distribution for a certain circular cylindrical bar

x1 + x2 T12 (x1 x2) 0 4 T] = T12 (x1 x2 ) x1 ; 2x2 0 0 0 x2

3 5

103

0 ; x3 T ] = 4 ; x3 0 x2 0
where is a constant.

x2 0 0

3 5

2 (a) What is the distribution of the stress vector on the surfaces dened by x2 2 + x3
and x1

= 4 x1 = 0,

= `? = `.

Find the total resultant force and moment on the end face x1 4.3 Nominal Stress Tensor

Solid mechanics problems are easy to formulate in the referential description. However the balance laws (4.1.14) and (4.1.15) derived earlier are in the spatial description. We note that the stressvector f in equations (4.1.5) and (4.1.6) is acting on a unit area in the present conguration of the
87

body. If we can express this f in terms of the area in the reference conguration, we would then be able to write the balance laws in the referential description. Note that

fida = Tij nj da = Tij daj = Tij J (F ;1)Bj dAB SiB = JTij (F ;1)Bj
equation (4.3.1) becomes

(4.3.1)

where we have substituted for dai in terms of dAB from eqn. (3.10.1). With the following denition (4.3.2)

fida = SiB dAB = SiB NB dA:

(4.3.3)

Thus SiB provides a measure of the present force acting on a unit area in the reference conguration. For this reason, SiB is called the nominal stress tensor, engineering stress tensor, or rst Piola-Kirchhoff stress tensor. Note that SiB NB equals the present

value of the force acting on the area in the present conguration at perpendicular to N at P is deformed.

P 0 into which a unit area

The stress tensors SiB and Tij are related by eqn. (4.3.2). However, to nd one stress tensor from the other, one must know the deformation gradient. In the Mechanics of Materials course, the nominal stress tensor SiB and the engineering strain eAB are used to plot the stress strain curve. In general, the graph of the true stress T11 versus the true strain 11 will look quite different from that of S11 versus e11 in a simple tension test in which the load is applied in the X1 -direction, and x1 and X1 axes point in the same direction. Substituting for fi da from (4.3.3) into (4.1.5) and writing dm = 0 dV , we obtain
Z Z Z

0 ai dV

SiB NB dA +
88

0 gi dV:

(4.3.4)

Now using the divergence theorem, we convert the surface integral on the right-hand side of (4.3.4) into the volume integral and thereby obtain
Z

( 0ai ; SiB B ; 0 gi)dV = 0:

(4.3.5)

This equation is to hold for all innitesimal volume elements in the body. If the integrand is continuous which is assumed to be the case, then (4.3.5) can hold for every volume element in the body if and only if

0 ai
or

= SiB B + 0 gi
iB = @S @X + 0 gi: B

(4.3.6)

0 ai

This is Cauchys rst law of motion in the referential description. From (4.1.15) and (4.3.2) we obtain the following for the second law of motion.

S] F]T = F] S]T :
Thus the nominal stress tensor SiB need not be and, in general, is not symmetric. The traction type boundary conditions in terms of SiB are written as

(4.3.7)

SiB NB = f^i
measured per unit area in the reference conguration. The following stress tensor is introduced in the engineering literature:

(4.3.8)

^i is the force acting at points on the boundary in the deformed conguration but is in which f

;1 T (F ;1 ) ~AB = JFAi S ij Bj
or

~ ] = J F;1 ] T] F;1 ]T = F;1 ] S]: S

(4.3.9)

~. is equivalent to the Cauchys second law of motion. There is no easy physical interpretation of S
89

~ known as the second Piola-Kirchhoff stress tensor is symmetric; the symmetry of S ~ The tensor S

4.4 Transformation of Stress Tensor under Rotation of Axes

Let us consider two sets of co-ordinate axes xi and x0i related by eqn. (2.7.4), that is,

e0i = aij ej :
normal n to this plane can be written as

(2.7.4)

Consider a plane the outer normal to which points in the positive x01 direction. Therefore, a unit

n = e01 = a1j ej
or

(4.4.1)

ni = a1i :
The stress vector f 0 on this plane is given by

(4.4.2)

fi0 = Tij nj = Tij a1j :


Recalling the interpretations of various components T11

(4.4.3) of the stress tensor given after

T12 etc.

0 eqn. (4.1.3), we note that T11

0 T 0 equal respectively, the components of f 0 in the positive T21 31

90

x01 x02 , and x03 directions. Thus


0 = f 0 e0 = a T a T11 1i ij 1j 1 0 = f 0 e0 = a T a T21 2i ij 1j 2 0 = f 0 e0 = a T a : T31 3i ij 1j 3
These three equations can be written as (4.4.4)

Tk0 1 = akiTij a1j :


Similarly, now consider planes whose outer normals point in the positive x02 proceed the way it was done to arrive at (4.4.5). The result is

(4.4.5)

x03 directions and


(4.4.6) (4.4.7)

Tk0 2 = akiTij a2j Tk0 3 = akiTij a3j :


Equations (4.4.5), (4.4.6) and (4.4.7) can be collectively written as

0 =a T a 0 T Tkp ki ij pj T ] = a] T] a] :

(4.4.8)

A comparison of this equation with eqn. (2.8.4) reveals that T is indeed a second order tensor. An application of eqn. (4.4.8) is the problem of nding the normal and the shear stress on an oblique plane when the stress-state on horizontal and vertical planes in the plane-stress problem is known.

91

For the axes shown,

cos 4 a] = ; sin 0
Therefore
2

sin cos 0

0 0 5: 1
32

0 T0 T0 T11 cos 12 13 0 T 0 T 0 5 = 4 ; sin 4 T21 22 23 0 T0 T0 T31 0 32 33


2

sin cos 0

0 0 1

32

T11 T12 0 cos 5 4 T12 T22 0 5 4 sin 0 0 0 0

; sin 0
cos 0 0 1

3 5 3

T11 cos2 + T22 sin2 + 2T12 sin cos 0 2 2 2 2 4 = T12 (cos ; sin ) + (T22 ; T11 ) sin cos T11 sin + T22 cos ; 2T12 sin cos 0 5 0 0 0 0 T 0 , the normal and shear stresses on the oblique plane. Equations relating Thus one can nd T11 12 0 T 0 to T11 T12 are called stress transformation equations in the Mechanics of Deformable T11 12
Bodies course. Exercise: The stress matrix

T referred to x1 x2x3 -axes is shown below at the left (in ksi).

New

x01 x02 x03 -axes are chosen by rotating the unprimed axes as shown to the right. x01 x02 x03 2 3 50 37:5 0 x1 0:6 0 0:8 T] = 4 37:5 ;25 ;50 5 x2 0 1 0 0 ;50 25 x3 ;0:8 0 0:6
a) Determine the traction vectors on each of the new coordinate ponents referred to the old (xi ) axes.

f = (;)e1 + (;)e2 + (;)e3 .

(x0i) planes in terms of comFor example, determine f on the x01 plane in the form

(b) Now project each of the vectors obtained in (a) onto the three new coordinate axes, and verify that the nine new components thus obtained for the stress matrix

T0 are the same as those

given by the formulae for the transformation of a second-order tensor under the rotation of axes. Exercise: The work given by

done by external forces during the deformation of a continuous body is


Z Z

W=

f uda +
92

g udv

By substituting for f from eqn. (4.1.12), using the divergence theorem and assuming that the body is deformed quasi-statically (that is, the inertia force
Z

ai is negligible), show that

W=
Hence show that for small deformations

Tij ui j dv: Tij eij dv:


Z

W= P=
By substituting for motion, show that
Z

Exercise: The power P of external forces is dened as

f vda +

g vdv:

f from eqn. (4.1.12), using the divergence theorem and Cauchys rst law of
_+ P =K
Z

Tij Dij dv

where
Z 1 K = 2 vi vidv

is the kinetic energy of the body. 4.5 Principal Stresses. Maximum Shear Stress In stress analysis problems one is interested in nding the maximum and the minimum normal stress and the maximum shearing stress at a point. In the following discussion, the point x in the present conguration is kept xed. is n. The stress vector f on this plane given by equation (4.1.12) is Tij nj . The normal stress on this plane is given by At a point x in the present conguration of the body, consider a plane whose outer unit normal

Tnormal = n f = niTij nj :
to nding the unit vector

(4.5.1)

Thus the problem of nding the maximum or the minimum normal stress at the point x reduces

n for which Tnormal is maximum or minimum.


93

Using the method of

Lagrange multiplier, the problem becomes that of nding extreme values of the function

F (n) = niTij nj ; (ni ni ; 1)


in which value is given by

(4.5.2)

is an arbitrary scalar. The vector n which makes F assume a maximum or minimum

@F = 0 @F = 0: @ni @
That is

(4.5.3)

(Tij ;

ij )nj

= 0 ni ni ; 1 = 0:
ij ] = 0

(4.5.4)

A nontrivial solution of eqn. (4.5.4)1 exists if and only if

det Tij ;
which gives the cubic equation

(4.5.5)

3 ; I 2 + II T T
Here IT

; IIIT = 0:

(4.5.6)

IIT and IIIT are the principal invariants of Tij . Since Tij is symmetric, therefore, eqn.
(2) (3) . Corresponding to each of these eigenvalues we

(4.5.6) has three real eigenvalues (1)

can nd a unit vector n from eqn. (4.5.4). For example, n(1) corresponding to (1) is given by

(Tij ;

(1)

(1) ij )nj

(1) = 0 n(1) j nj = 1:

(4.5.7)

n(1) n(2) n(3) are normals to planes on which Tnormal assumes extreme values. Since
(1) (1) (1) Tnormal = n(1) i Tij nj = ni
therefore (1) plane with outer unit normal n(1) is given by

(1)

(1) ij nj

(1)

(4.5.8)

(2) and (3) are extreme values of normal stresses. Also the stress vector on the

fi = Tij n(1) j =
Let

(1)

(1) ij nj

(1) n(1) : i

(4.5.9)

e a unit vector in this plane.

Then the shear stress in the direction of

e on the n(1) plane is


(4.5.10)

given by

Tshear = f e =
94

(1) e n(1) i i

= 0:

The plane on which the shear stress is zero is called a principal plane and the normal stress on a principal plane is called a principal stress. Thus (1)

(2) and (3) are principal stresses and

n(1) n(2) and n(3) are normals to principal planes. n(1) n(2) and n(3) are called principal axes of
the stress.

n(1) n(2) and n(3) are uniquely determined and are mutually orthogonal. However, when any two or all three of the principal stresses are equal, then n(1) n(2) and n(3) are not uniquely determined but can still be found so that they are mutually orthogonal.
Whenever (1)

6=

(2)

6=

(3)

Henceforth, in this section, we will assume that the principal axes of the stress tensor are mutually orthogonal. Taking x1

x2 x3 axes along the principal axes of the stress, we see that with respect to these
2

axes Tij has the form

Tij ] = 4 0 0

(1)

0 0

(2)

0 0

3 5:

(3)

(4.5.11)

We now nd the plane of the maximum shear stress. Let an outer unit normal to this plane be

n. Then, taking principal axes of stress as the coordinate axes,


fi = Tij nj =
(1) n 1 i1 + (2) n (3) n 2 i2 + 3 i3
(4.5.12)

gives the traction on this plane. Therefore the shear stress on this plane is given by

(Tshear )2 = fi fi ; (fini)2 =
(1)2 n2 + (2)2 n2 + (3)2 n2 ; ( (1) n2 + (2) n2 + (3) n2 )2 : 1 2 3 1 2 3
(4.5.13) Thus the problem of determining maximum Tshear reduces to that of nding extreme values of the function

2 + (n n ; 1) G(n) = Tshear i i
in which is a Lagrange multiplier. The necessary conditions are

(4.5.14)

@G = 0 @G = 0 @ni @
95

which are equivalent to

(i)2 n

i;(

(1) n2 + (2) n2 + (3) n2 )(2 (i) )n + i 1 2 3 i

ni = 0 (no summation on i)

(4.5.15) (4.5.16)

nini = 1 :
Three equations given by (4.5.15) for i

= 1 2 and 3 and the eqn. (4.5.16) determine ni and . A solution of these equations for which two out of n1 n2 and n3 are zero is not interesting since this corresponds to nding Tshear on a principal plane which is zero. The solution of (4.5.15) and (4.5.16) for which n1 6= n2 6= n3 6= 0 exists only if (1) = (2) = (3) . In this case too, eqn. (4.5.13) gives Tshear = 0. The remaining possibility is that one out of n1 n2 and n3 be zero, say n1 = 0, n2 6= 0 6= n3 . For this case, eqn. (4.5.15) for i = 1 is satised identically. For i = 2 and 3 we obtain
(2)2 n ; ( (2) n2 + (3) n2 )2 (2) n + 2 2 2 3 (3)2 n ; ( (2) n2 + (3) n2 )2 (3) n + 3 3 2 3
and, of course,

n2 = 0 n3 = 0

2 n2 2 + n3 = 1:
A solution of these is

1 : n2 = n3 = p 2 For the values of n2 and n3 given by (4.5.17) and n1 = 0, eqn. (4.5.13) gives 2 = 1 ( (2) ; (3) )2 Tshear 4
and hence

(4.5.17)

Tshear = 1 2(
Similarly for n2 this plane is

(2) ; (3) ):

(4.5.18)

1 and the shear stress on = 0, a solution of (4.5.15) and (4.5.16) is n1 = n3 = p 2

Tshear = 1 2(
96

(1) ; (3) ):

(4.5.19)

For n3 is

1 and the shear stress on this plane = 0, a solution of (4.5.15) and (4.5.16) is n1 = n2 = p 2

Tshear = 1 2(

(1) ; (2) ):

(4.5.20)

Thus the extreme values of the shear stress are given by (4.5.18), (4.5.19) and (4.5.20). Note that on planes of maximum shear stress, the normal stress need not be zero. In fact, the normal stress on the plane for which

1 p 1 n= 0 p 2

1( is 2

(2) + (3) ):

5 The Linear Elastic Material


5.1 Introduction So far we have studied the kinematics of deformation, the description of the state of stress and three basic conservation laws of continuum mechanics, viz., the conservation of mass (eqn. 3.11.8), the conservation of linear momentum (eqn. 4.1.14), and the conservation of moment of momentum (eqn. 4.1.15). All these relations are valid for every continuum, indeed no mention was made of any material in their derivations. These equations are, however, not sufcient to describe the response of a body to a given loading. We know from experience that under the same loading conditions, identical specimens of steel and rubber deform differently. Furthermore, for a given body, the deformations vary with the loading conditions. For example, for moderate loadings, the deformation in steel caused by the application of loads disappears with the removal of loads. This aspect of the material behavior is known as elasticity. Beyond a certain loading, there will be permanent deformations, or even fracture exhibiting behavior quite different from that of elasticity. In this chapter, we shall study an idealized linear elastic material for which the stress-strain relationship is linear. Using this stress-strain law, we will then study some dynamic and static problems. Henceforth in this chapter we will assume that the deformations are small so that eAB is an adequate measure of strain.
97

5.2 Linear Elastic Solid. Hookean Material For a linear elastic solid or a Hookean material, it is assumed that the Cauchy stress is a linear function of the innitesimal strain. That is,

T11 = C1111 e11 + C1112 e12 + T12 = C1211 e11 + C1212 e12 +

C1133 e33 C1233 e33


(5.2.1)

::::::::::::::::::::::::::::

T33 = C3311 e11 + C3312 e12 + C3333 e33 : We will assume that Tij is symmetric and since eAB is also symmetric, the above six equations can
be written as

Tij = Cijklekl
in which

(5.2.2)

Cijkl = Cjikl = Cijlk :

(5.2.3)

Since Tij and eij are components of second order tensors, Cijkl are components of a fourth-order tensor. It is known as the elasticity tensor. It is this tensor which characterizes the mechanical properties of a particular anisotropic Hookean elastic solid. The anisotropy of the material is represented by the fact that the components of Cijkl are in general different for different choices of coordinate axes. If the body is homogeneous, that is, the mechanical properties are the same for every particle of the body, then Cijkl are constants (i.e. independent of position). We shall only study homogeneous bodies. Because of the symmetry relations (5.2.3), the fourth order tensor

Cijkl has 36 independent

components. Thus, for a linear anisotropic elastic material, we need no more than 36 material constants to specify its mechanical properties. It follows from (5.2.1) that whenever eij

= 0 Tij = 0.

Thus, in the reference conguration,

there is no stress. This implies that there are no initial stresses present. A material is said to be isotropic if its mechanical properties can be described without reference to direction. That is, the components of the elasticity tensor Cijkl remain the same regardless of
98

how the rectangular Cartesian coordinate axes are rotated. In other words,

0 Cijkl = Cijkl

(5.2.4)

under all orthogonal transformations of coordinate axes. A tensor having the same components with respect to every orthonormal basis is known as an isotropic tensor. An example of an isotropic tensor is ij . It is obvious that the following three fourth-order tensors are isotropic:

Aijkl =

ij kl

Bijkl =

ik jl

Dijkl =

il jk :

In fact, it can be shown that any fourth order isotropic tensor can be represented as a linear combination of the above three tensors. Thus, for an isotropic, linear elastic material, the stress-strain law (5.2.2) can be written as

Tij = ( Aijkl + Bijkl + Dijkl)ekl:


In order for the symmetry relation (5.2.3)2 to hold,

. Thus

Tij = (

ij kl +

ik jl +

il jk )ekl
(5.2.5)

= ekk ij + 2 eij :
In the preceding equations and

are constants. Equation (5.2.5) is the constitutive equation and are known as Lames

for a linear elastic isotropic material. The two material constants constants. Since eij are dimensionless, and

are of the same dimensions as the stress tensor,

force per unit area. For a given material, the values of Lam es constants are to be determined from suitable experiments. We now write (5.2.5) in a form usually studied in the Mechanics of Deformable Bodies course. Taking the trace of both sides of (5.2.5) we obtain

Tii = (3 + 2 )ekk :
Assuming that

(5.2.6)

(3 + 2 ) 6= 0
99

6= 0

(5.2.7)

we get

1 T ekk = (3 + 2 ) kk
and hence

eij = 21 Tij ; 3 + 2 Tkk

ij

(5.2.8)

To get a physical interpretation of Lam es constants in terms of Youngs modulus E and Poissons ratio , we note that, in a simple tension test, with the axial load P applied along X1 -axis,

Tij = P A
11 e11 = T E

i1 j 1

where A is the area of cross-section of the prismatic body. Also

= ; e22 : e11 = 2( + ) :

(5.2.9)

Substituting for e11 and e22 from (5.2.8) into (5.2.9) and assuming that (

+ ) 6= 0 we arrive at
(5.2.10)

+2 ) E = (3 +
The elimination of from these two equations gives

=
that E must be > 0.

E : 2(1 + )

(5.2.11)

In order for a tensile load to produce extension in the direction of loading, it is clear from (5.2.9)1

Another stress state, called simple shear, is the one for which all stress components except one pair of off-diagonal elements vanish. In particular, we choose T12 gives

= T21 6= 0.

Equation (5.2.8)

12 e12 = T 2 :
of the applied load, e12 and T12 must be of the same sign. For that to be true,

(5.2.12)

In order for a simple shear force on X2 plane to produce a sliding of the X2 plane in the direction must be > 0.

100

For E and

to be positive, it follows from (5.2.11) that ;1

<

. Now consider a stress state

called, hydrostatic stress, for which Tij

= ;p

ij . For this case, equation (5.2.8) gives


(5.2.13)

Since ekk

;dV , therefore, in order for the hydrostatic pressure to produce a decrease in volume, = dvdV

ekk = ; ; 1 2 p : +3

we must have

+ (2=3) > 0 :
Substituting in this equation from (5.2.10), we obtain

(5.2.14)

2 3
On the assumption that Rewriting (5.2.13) as

1+ 1;2

>0:
must be

(5.2.15)

> 0, > ;1, the inequality (5.2.15) implies that


2 p ekk = ; 23 11; +

1=2.

(5.2.16)

= 1=2 ekk = 0. That is, there is no change in volume. Since for incompressible materials ekk = 0, therefore, = 1=2 for incompressible materials.
we see that for Equations (5.2.10), (5.2.11) and (5.2.5) yield

Tij = 1 2 ; 2 ekk ij + 2 eij :


and hence is indeterminate. It is usually denoted by incompressible linear elastic material becomes

(5.2.17)

For an incompressible material, the rst term on the right-hand side of (5.2.17) is of the form %

;p ij and the constitutive relation for an


(5.2.18)

Tij = ;p ij + 2 eij :

Here p is called the hydrostatic pressure, and it can not be determined from the strain eld. However, whenever surface tractions are prescribed on at least a part of the boundary, p can be uniquely determined. Example:
101

a) For an isotropic Hookean material, show that the principal axes of stress and strain coincide. b) Find a relation between the principal values of stress and strain. Solution: Note that eigenvectors of Tij and eij are the principal axes of stress and strain respectively. Let n(1) be an eigenvector of eij and e(1) be the corresponding eigenvalue. That is

(1) (1) eij n(1) j = e ni :


By Hookes law we have

(1) Tij n(1) j = ( ekk ij + 2 eij )nj (1) = ekk n(1) i + 2 eij nj (1) (1) = ekk n(1) i + 2 e ni

= ( ekk + 2 e(1) )n(1) i :


Therefore, n(1) is also an eigenvector of Tij and the corresponding eigenvalue is

( )

b) It is clear from eqn. ( ) that the eigenvalue T (1) of Tij corresponding to the eigenvector n(1) is ekk + 2 e(1) . Since ekk = e(1) + e(2) + e(3) , therefore,

ekk + 2 e(1) .

T (1) = (e(1) + e(2) + e(3) ) + 2 e(1) :


Similarly,

T (2) = (e(1) + e(2) + e(3) ) + 2 e(2) T (3) = (e(1) + e(2) + e(3) ) + 2 e(3) :
Exercise: (Recall the exercise given on page 2-11). For W

Tij .
Substitute e21

@W = = eij eij + 2 (ekk )2, show that @e ij

Hint: It is advisable to do the problem long-hand. That is, rst expand the given expression for W .

= e12 = (e12 + e21 )=2 etc. in it and then carry out the differentiation with respect to

e11 e22 e12 etc. W is called the stored energy function or the strain energy density.
102

5.3 Equations of the Innitesimal Theory of Elasticity We shall consider only the case of small strains, and innitesimal velocities and accelerations as compared to some reference values. Thus every particle is always in a small neighborhood of the reference conguration. The reference conguration in which Tij

= 0 is also called a natural state.

Thus, if xi denotes the position in the natural state of a typical material particle, we assume that @ui is much xi ' Xi, and that the magnitude of the components of the displacement gradient @X j smaller than unity. Since

xi = Xi + ui
we have

@Tij = @Tij @xk @X1 @xk @X1 @u1 + @Tij @u2 + @Tij @u3 ij 1 + = @T @x1 @X1 @x2 @X1 @x3 @X1 ij ' @T @x1 :
Similarly,

@Tij @X2
Therefore, for small deformations,

@Tij @Tij @x2 @X3 @Tij @Xk @Tij : @xk

@Tij : @x3
(5.3.1)

Since

2 i ai = @@tx 2
and

Xj ; xed

2 i = @@tu 2

Xj ; xed

= 0 (1 ; ekk )
we have

2 i ai = 0 (1 ; ekk ) @@tu 2

@ 2 ui 0 2 @t

Xj ; xed

(5.3.2)

:
103

Also

gi ' 0gi :
Thus the balance laws or conservation laws for small deformations of a body take the following form.

= 0 (1 ; ekk ) @ 2 ui @Tij : Balance of linear momentum: 0 2 = 0 gi + @t @Xj


Balance of mass: (5.3.4) gives

(5.3.3) (5.3.4)

For a linear elastic homogeneous and isotropic body, substitution from (5.2.5) and (3.14.3) into

@ 2 ui = g + ( + )u + u : 0 2 0 i k ki i jj @t
These are three equations for the three unknowns u1

(5.3.5) After a solution of (5.3.5)

u2, and u3.

has been obtained, one can nd the present mass density from eqn. (5.3.3). Note that eqn. (5.3.5) is a system of three coupled partial differential equations. In order to nd a solution of (5.3.5) applicable to a given problem, side conditions such as initial conditions and boundary conditions are needed. In a dynamic problem, one needs the values of ui (Xj

0) and u _ i(Xj 0).

That is, the initial

displacement and the initial velocity eld should be given as smooth functions throughout the body. Note that these initial conditions are not needed in a static problem. However, in both static and dynamic problems one needs boundary conditions which can be one of the following three types. In the boundary condition of traction the stress vector is prescribed at the boundary points of the body. That is, at the points on the boundary

Tij nj = uk kni + (ui j + uj i)nj = fi(X t)


in which fi is a known function.

(5.3.6)

In a displacement type boundary condition, displacements are prescribed on the boundary points. For example, a part of the boundary of a body could be glued to a rigid support. In this case,
104

displacements for these boundary points will be zero. The third type of boundary condition is the one in which surface tractions are prescribed on one part and the displacements on the remainder or at a boundary point, tangential components of the stress-vector and the normal component of the displacement vector (or vice-versa) are prescribed. These are known as the mixed type boundary conditions. 5.4 Principle of Superposition

(1) (2) (1) Let ui and ui be two possible displacement elds corresponding to the body force elds gi (2) (1) (2) (1) (2) and gi and surface tractions fi and fi respectively. Let Tij and Tij be the stress elds (1) (2) corresponding to displacement elds ui and ui . Then @ 2 u(1) (1) (1) (1) i (5.4.1) 0 @t2 = 0 gi + ( + )uk ki + ui jj @ 2 u(2) i = g (2) + ( + )u(2) + u(2) (5.4.2) 0 i 0 i jj k ki 2 @t (1) (1) (1) (1) on the boundary (5.4.3) either ui = hi (Xj t) or Tij nj = fi (2) (2) (2) either u(2) (5.4.4) i = hi (Xj t) or Tij nj = fi on the boundary:
Adding (5.4.1) to (5.4.2), and (5.4.3) to (5.4.4) we obtain

(2) @ 2 (u(1) i + ui ) = (g (1) + g (2) ) + ( + )(u(1) + u(2) ) + (u(1) + u(2) ) 0 i 0 i i i jj k k ki @t2 (1) (2) (1) (2) (1) (2) (1) (2) either ui + ui = hi + hi or (Tij + Tij )nj = fi + fi on the boundary: (1) (2) Thus ui + ui is a possible motion for the same linear elastic body corresponding to the body (1) (2) (1) (2) force gi + gi and surface tractions fi + fi . This is the principle of superposition and is
frequently used in the Mechanics of Materials course when solving problems for the combined loads. One application of this principle in linear elastic problems is that in a given problem, we shall often assume that the body force is absent having in mind that its effect, if not negligible, can always be obtained separately and then superposed onto the solution of the problem with vanishing body force.

105

5.5 A Uniqueness Theorem Equations (5.3.5) subject to given initial and boundary conditions have a unique solution. (1) (2) Assume that there exist two solutions ui and ui of eqn. (5.3.5) subject to the same initial and boundary conditions. Then

(2) wi = u(1) i ; ui will be a solution of @ 2 wi = ( + )w + w 0 2 k ki i jj @t


and, on the boundary, either

(5.5.1)

Tij nj = wk k ni + (wi j + wj i)nj = 0


or

(5.5.2)

wi = 0
and

(5.5.3)

wi(Xj 0) = 0 w _ i(Xj 0) = 0:
Taking the inner product of (5.5.1) with w _ i(Xj

(5.5.4)

t), integrating the resulting equation over the

region R occupied by the body in the reference conguration and by using the divergence theorem, we arrive at
Z Z d Tij nj w _ idA ; Tij w _ j idV = dt Z

w _ iw _ i dV: 2

(5.5.5)

The rst integral vanishes because of (5.5.2) and (5.5.3). The second integral can be rearranged to read

d Z w w + e e dV ij ij dt 2 k k i i
in which eij

= (wi j + wj i)=2. Thus eqn. (5.5.5) becomes

d Z e e + e e + 0w _ i dV = 0: ij ij dt 2 kk ii 2 _ iw
106

(5.5.6)

Integrating this and making use of the initial conditions (5.5.4) for wi , we obtain
Z

e e + eij eij + 0 w _w _ dV = 0: kk ii 2 2 i i
de d + 0 w + e e + e _i kk ii ij ij 2 3 2 _ iw

(5.5.7)

Noting that the integrand can be written as (5.5.8)

in which eijd

ij = eij ; ekk 3 , we see that every term in (5.5.8) is positive provided that

3 +2 >0
We will henceforth assume that and

> 0:

(5.5.9)

satisfy (5.5.9). Thus for eqn. (5.5.7) to hold, (5.5.10)

eijd = 0 eii = 0 w _ i = 0:
Since wi

(2) = 0 initially, therefore, wi(Xj t) = 0 which implies that u(1) i (Xj t) = ui (Xj t).

Thus if somehow one can nd a solution of eqn. (5.3.5) that satises the prescribed initial and boundary conditions, then that is the only solution of eqn. (5.3.5). There are very few dynamic problems that have been solved. In a static problem or, more appropriately, in a quasi-static problem, the left-hand side of eqn. (5.3.5) becomes zero and one needs only the prescribed boundary conditions. Thus the difference (1) (2) solution wi = ui ; ui will satisfy

0 = ( + )wk ki + wi jj
grating the resulting equation over R, and using the divergence theorem we arrive at
Z

(5.5.11)

and either (5.5.2) or (5.5.3) on the boundary. Taking the inner product of (5.5.11) with wi , inteZ

Tij nj widA ; Tij wj idV = 0:

(5.5.12)

The rst integral vanishes because of (5.5.2) and (5.5.3) and the second integral can be written as
Z

d d 2 + 3 ekk eii + eij eij dV = 0:

(5.5.13)

107

For this equation to hold,

eijd = 0 eii = 0
and hence

eij = 0:
A solution of eqn. (5.5.14) given in Section 3.14 is

(5.5.14)

wi = ai + bij Xj
in which ai and bij

(5.5.15)

= ;bji are constants. If displacements are prescribed at three noncolinear points on the boundary, then wi = 0 and the solution of the given boundary value problem is
unique. However, if surface tractions are prescribed on all of the boundary, then different solutions of the same boundary value problem can differ at most by a rigid body motion. Even though the two displacement elds differ by a rigid body motion, the stress elds and the strain elds obtained from such displacement elds are identical. 5.6 Compatibility Equations Expressed in terms of the Stress Components for an Isotropic, Homogeneous, Linear Elastic Solid In static problems where the surface tractions are prescribed on the entire boundary, it is convenient to solve the equations of equilibrium

Tij j + 0gi = 0

(5.6.1)

in terms of the stress components. Having obtained Tij which satisfy (5.6.1) and the assigned boundary conditions, we solve for strains from eqn. (5.2.8). In order that eij give a unique displacement eld ui

eij

must satisfy the compatibility conditions (3.14.32). Substitution for eij

from (5.2.8) into (3.14.32) yields

1 T +T ;T ;T ] 2 ij kp kj ij ip jk jk ip ; 2 (3 + 2 ) ij Tmm kp + Tmm ij

kp ; Tmm ip jk ; Tmm jk ip ] = 0:

(5.6.2)

108

There are only 6 independent equations as there were only 6 independent equations expressed by (3.14.32). In (5.6.2) set k

= p and sum with respect to the common index to obtain


(5.6.3)

Tij kk + Tkk ij ; Tik jk ; Tjk ik = 3 + 2 Tmm kk ij + Tmm ij ]:


Out of these 9 equations only six are independent because Tij

= Tji. Consequently, in combining

linearly some of the equations (5.6.2), the number of independent equations has not been reduced, and hence the resulting set (5.6.3) of equations is equivalent to the original set (5.6.2). In solid mechanics problems, it is usual to neglect the effect of gravity and write (5.6.1) as

Tij j = 0:

(5.6.4)

Whenever (5.6.4) holds, the third and fourth terms on the left hand side of (5.6.3) vanish and it simplies to

+ )T ; Tij kk + 2( 3 + 2 kk ij 3 + 2 Tmm kk ij = 0:
Setting i to

(5.6.5)

= j and summing over the repeated index, we get Tmm kk = 0 and hence (5.6.5) reduces + ) T = 0: Tij kk + 2( 3 + 2 kk ij

(5.6.6)

These are the compatibility equations in the absence of body forces. Thus for a stress eld to be a solution of a static problem with zero body force for a homogeneous and isotropic linear elastic body, it must satisfy (a) equations of equilibrium (5.6.4), (b) compatibility conditions (5.6.6), and (c) the appropriate boundary conditions. Note that if Tij satises (a) and (c) but not (b), then that Tij will not correspond to a stress eld in a linear elastic body since eij corresponding to such a Tij will not result in a unique ui .

109

Exercise: Can the following stress eld represent a possible solution of a static problem with zero body force for a homogeneous and isotropic linear elastic body?

T11 = c(X22 + (X12 ; X22)) c = constant 6= 0 T22 = c(X12 + (X22 ; X12)) T33 = c (X12 + X22) T12 = ; 2c X1X2 T23 = T31 = 0:
Example: For a plane stress state, express the compatibility conditions in terms of a stress function. Solution: Recall the second exercise given on page 4-9. For a plane stress state,

T11 =

22

T22 =

11

T12 = ;

12

= (X1 X2) T13 = T23 = T33 = 0:


As was proved in that exercise, the stress eld thus obtained from compatibility condition. This we obtain from (5.6.6) by setting i peated index. The result is

(5.6.7)

satises the equilibrium equa-

tions (5.6.4) for every choice of the function . For this state of stress there is only one independent

= j and summing over the re(5.6.8)

Tii kk = 0:
Now substitution from (5.6.7) into (5.6.8) gives

(
or

22 + 11 ) 11 + ( 22 +

11 ) 22

=0

r2 (r2 ) = r4 = 0:
under the appropriate boundary conditions. Exercise: Show that for plane strain problems (i.e., those for which u1

(5.6.9)

Thus the problem of solving a static plane stress problem reduces to nding a solution of (5.6.9)

= u1(X1 X2) u2 =

u2(X1 X2) u3 = 0) with zero body force there is only one independent compatibility condition in
terms of the components of the stress and this can be written as

r2 (T11 + T22 ) = 0:
110

(5.6.10)

Exercise: (a) Simplify the equations of equilibrium with zero body force for plane strain problems. (b) Introduce a function

(X1 X2) such that T11 =

22

T22 =

11

T12 = ;

12 :

(5.6.11)

Will this stress distribution satisfy the equilibrium equations obtained in part (a)? Note that for plane strain case T33 need not and, in general, will not vanish. If we substitute the stress eld given by (5.6.11) into (5.6.10) we arrive at (5.6.9). Thus the task of nding a solution of a static plane strain problem also reduces to that of nding a solution of (5.6.9) which satises the pertinent boundary conditions. 5.7 Some Examples (a) Vibration of an Innite Plate Consider an innite plate bounded by the planes

X1 = 0 and X1 = `.

The vibrations of the

plate are caused by a prescribed motion of these bounding planes or by prescribed surface tractions on these planes. Since we study steady state vibration of the plate, the initial displacement and velocity elds in the plate are not required. We will neglect the effect of gravity in these problems. We begin by assuming that the displacement eld in the plate is of the form

u1 = u1(X1 t) u2 = u3 = 0: @ 2 u1 = ( + 2 ) @ 2 u1 @t2 @X12

(5.7.1)

For this displacement eld, the equations of motion (5.3.5) reduce to the following equation

0
which can be rewritten as

(5.7.2)

@ 2 u1 = C 2 @ 2 u1 C 2 = + 2 : L @X 2 L @t2 0 1
A steady state vibration solution to this equation is of the form

(5.7.3)

u1 = (A cos X1 + B sin X1)(C cos CL t + D sin CL t)


where the constants

(5.7.4)

A B C D and

are determined by boundary conditions. This vibration

mode is sometimes termed a thickness stretch because the plate is being stretched through its
111

thickness. It is analogous to acoustic vibration of organ pipes and to the longitudinal vibration of slender rods. Another vibration mode can be obtained by assuming the displacement eld

u2 = u2(X1 t) u1 = u3 = 0:
In this case, the displacement eld must satisfy the equation

(5.7.5)

2 2 2 @ u2 = @ u2 C 2 = CT @X 2 @t2 T 1
and the solution is of the same form as in the previous case.

(5.7.6)

This vibration is termed thickness shear and it is analogous to the motion of a vibrating string. Example 5.7.1: (a) Find the thickness-stretch vibration of a plate, where the left face (X1 subjected to a forced displacement ui

= 0) is

= ( cos !t)

i1 and the right face (X1

= `) is xed.

(b) Determine the values of ! that give resonance. Solution: (a) The boundary condition on the left face X1

= 0 and eqn. (5.7.4) give

cos !t = A(C cos CL t + D sin CL t):


Therefore,

AC =
The second boundary condition gives

! D = 0: =C
L

0 = u1(` t) =
Therefore,

!` + BC sin !` cos !t: cos C C


L L

!` BC = ; cot C L
and the vibration is given by
3 ! 6 ! X ; sin C1 X1 7 7 u1(X1 t) = 6 cos 4 CL 1 tan !` 5 cos !t: C 2

112

(b) Resonance is indicated by unbounded displacements. This occurs in part (a) for forcing frequencies corresponding to tan

!` = 0, that is, when CL ! = n `CL n = 1 2 3 : : :


i2 on the plane X1

Example 5.7.2 (a) Find the thickness-shear vibration of an innite plate which has an applied surface traction fi

= ; cos !t

= 0 and is xed on the plane X1 = `.

(b) Determine the resonant frequencies. Solution: On the plane X1

= 0, n = (;1 0 0). Therefore

fi = Tij nj = ;Ti1
gives

T21 X1 =0 = cos !t:


This shearing stress forces a vibration of the form

u2 = (A cos X1 + B sin X1 )(C cos CT t + D sin CT t):


Using Hookes law (5.2.5), we have

@u2 : T12 = 2 e12 = @X 1


Therefore

@u2 @X1 X1 =0 = cos !t (B )(C cos CT t + D sin CT t) = cos !t:


Thus

D=0
The boundary condition at X1

! BC = CT : =C ! T

= ` gives u2(` t) = 0. (A cos ` + B sin `)C cos CT t = 0:


113

Hence

A = ;B tan `:
Thus

CT sin ! X ; tan !` cos ! X cos !t: u2(X1 t) = ! CT 1 CT CT 1


(b) Resonance occurs for

!` = 1 tan C
T
or

T ! = n 2C `
(B) Torsion of a Circular Shaft

n = 1 3 5 :::

Consider elastic deformations of a cylindrical bar of circular cross-section of radius a and length

L, that is being twisted by an end moment Mt at the right end and is xed at the left end. We choose the X3 -axis to coincide with the axis of the cylinder and the left-hand and right-hand faces to correspond to the planes X3 = 0 and X3 = L respectively.

This problem involves the solution of equilibrium equations

( + )uj ji + ui jj = 0

(5.7.7)

114

subject to the boundary conditions

Tij nj = 0 on the lateral surface, R Tij nj dA = 0 on the plane X3 = L R "ijk Xj TkpnpdA = Mt i3 on the plane X3 = L ui = 0 on the plane X3 = 0:

9 > > > > > = > > > > >

(5.7.8)

Note that if the problem is formulated in terms of displacements, then compatibility conditions (5.6.6) are not needed. We use St. Venants semi-inverse method to solve the problem in an inverse way. That is, we make a kinematic assumption about the displacement eld and then ensure that eqns. (5.7.7) and (5.7.8) are satised. Because of the symmetry of the problem, it is reasonable to assume that the motion of each cross-sectional plane induced by the end moments is a rigid body motion about the X3 -axis. This motion is similar to that of a stack of coins in which each coin is rotated by a slightly different angle than the previous coin. We will see that this assumption results in a displacement eld that satises (5.7.7) and (5.7.8). To ensure that the deformations are small, we will assume that the angle of rotation of any section with respect to the left-end is very small as compared to one. Under the preceding assumptions, the displacement of any point can be calculated as follows.

u1 = r cos( + ) ; r cos ' ; X2 u2 = r sin( + ) ; r sin ' X1 ' ; X2 u3 = 0:


For this displacement eld to satisfy the equations of equilibrium (5.7.7), we must have

d2 = 0: dX32
Thus

d = = constant: dX3
115

(5.7.9)

That is, the angle of twist per unit length is the same over the entire length of the shaft. From (5.7.9) and recalling that

= 0 at X3 = 0, we get = X3

and, therefore,

u1 = ; X2 X3 u2 = X1X3 u3 = 0:

(5.7.10)

The strains and stresses associated with these displacements can be calculated from eqns. (3.14.3) and (5.2.5). The non-zero components of strains are

e23 = e32 = x1 =2 e31 = e13 = ; X2 =2


and the non-zero components of stress are

(5.7.11)

T13 = T31 = ; X2 T23 = T32 =


a unit outer normal n =

X1:

(5.7.12)

To see whether the boundary conditions are satised or not, we note that on the lateral surface

1 (X X 0). Therefore, on the lateral surface, a 1 2 X2 1 Tij nj = Ti1 X + T i2 a a

= L, n = (0 0 1) and R R R in order that (5.7.8)2 be satised, Ti3 dA = 0 which is true because X1 dA = 0, X2 dA = 0. The boundary condition (5.7.8)3 requires that, on the end plane X3 = L,
and the calculated stress eld does satisfy (5.7.8)1 . On the end plane X3
Z

"ijkXj Tk3dA = Mt i3 :
Z

This gives

(X12 + X22)dA = Mt

or

J p = Mt
116

(5.7.13)

in which Jp is the polar moment of inertia and equals clearly satised by the displacement eld (5.7.10). Substituting for

a4 =2.

The boundary condition (5.7.8)4 is

from (5.7.13) into (5.7.10), (5.7.11) and (5.7.12) we obtain the displace-

ment components, non-zero strain components and non-zero stress components at any point in the cylindrical bar. In terms of the twisting moment Mt , the stress tensor becomes
2 3

0 0 ;X2 6 7 t6 0 7 0 X Tij ] = M (5.7.14) 1 5 Jp 4 ;X2 X1 0 At any point (X1 X2 b) of a cross-section X3 = b n = (0 0 1) and the stress vector f is given

by

fi = Tij nj =
Note that f lies in the plane X3

(;X2 1i + X1 2i ) :

= b implying thereby that there is no normal stress at any point on the plane X3 = b. Also f is perpendicular to the radius vector joining the point (X1 X2 b) with the center (0 0 b) of the cross-section. The magnitude of f is q t jf j = r = M r r = X12 + X22 : Jp
From this we see that the maximum stress is a tangential stress that acts on the boundary of the Mt a. cylinder and has the magnitude Jp (c) Torsion of Non-Circular Cylinders For cross-sections other than circular, the stress eld (5.7.14) does not satisfy the boundary conditions of zero tractions on the mantle of the cylinder. We will see that in order for this boundary condition to be satised, the cross-sections will not remain plane. We begin by assuming a displacement eld that still rotates each cross-section by a small angle , but in addition there may be a displacement in the axial direction. This warping of a crosssectional plane will be dened by u3

= ^ (X1 X2).

Thus we assume that each cross-section is

warped in the same way. For this to be true the left end of the cylinder can not be xed to a rigid at support but is subjected to a torque equal and opposite to that applied to the right end. The rigid motion is removed by applying suitable constraints.
117

The assumed displacement eld has the form

u1 = ;X2 (X3 ) u2 = X1 (X3) u3 = ^(X1 X2):


For this displacement eld to satisfy the equations of equilibrium (5.7.7), we must have

(5.7.15)

d2 = 0 dX32 @ 2 ^ + @ 2 ^ = 0: @X12 @X22


Therefore, the angle Setting

(5.7.16)

of twist per unit length is the same over the entire length of the cylinder.

= 0 at X3 = 0, we get = X3 and therefore the displacement eld (5.7.15) can be

rewritten as

u1 = ; X2X3 u2 = X1X3 u3 =
where

(X1 X2)

(5.7.17)

^=

. For equilibrium,

must satisfy

@ 2 + @ 2 = r2 = 0: @X12 @X22
Stresses corresponding to the displacements (5.7.17) are

(5.7.18)

Since the bar is cylindrical, the unit normal to the lateral surface has the form n and the associated surface traction is given by

T11 = T22 = T33 = T12 = 0 @ +X T = T23 = 1 13 @X2

@ ;X : 2 @X1

(5.7.19)

= n1e1 + n2e2

fi =

@ ;X n + @ +X n 2 1 1 2 @X1 @X2
must satisfy

i3 :

For the lateral surface to be traction free,

d = @ n + @ n =X n ;X n 2 1 1 2 dn @X1 1 @X2 2
on the boundary. In order that the boundary condition (5.7.8)2 be satised,
Z Z

(5.7.20)

T13 dA = 0

T23 dA = 0:

118

Now
Z

T13 dA =
= = =0

Z Z I

@ ; X dA 2 @X1 @ X @ ;X + @ X @ +X 2 1 @X1 1 @X1 @X2 1 @X2 @ ; X n + @ + X n ds X1 @X 2 1 1 2 @X2 1


R

dA

because of (5.7.20). Using a similar argument one can show that condition (5.7.8)3 gives
Z

T23 dA = 0.

The boundary

Mt = (X1T23 ; X2T13 )dA


= =D
where
Z Z

@ ; X @ dA X12 + X22 + X1 @X 2 @X
2 1
(5.7.21)

D=

@ ; X @ dA: X12 + X22 + X1 @X 2 @X


2 1

(5.7.22)

The formula (5.7.21) shows that the twisting moment or torque Mt is proportional to the angle of twist per unit length, so that the constant D provides a measure of the rigidity of a bar subjected to torsion.

D is called the torsional rigidity of the bar.


is determined by solving (5.7.18) and (5.7.20).

Hence the torsion problem is solved once

For an elliptic cylindrical bar with the cross-section given by

X12 + X22 = 1 a2 b2
we see that on the boundary,

(5.7.23)

1 X1 e + X2 e n= F a2 1 b2 2
and, therefore, eqn. (5.7.20) becomes

2 2 1 + X2 F2 = X a4 b4

(5.7.24)

@ X b2 + @ X a2 = X X (b2 ; a2 ): 1 2 @X1 1 @X2 2


119

(5.7.25)

This suggests that

2 ; a2 =b b2 + a2 X1 X2:
This choice of satises (5.7.25) and (5.7.18). Substituting for

(5.7.26) in (5.7.22) and then the result in

(5.7.21) we arrive at

2 I + a2 I ] b Mt = b2 2 11 2 + a 22
in which I11 and I22 are, respectively, the second moments of area about X1 and X2 axes. Recalling that for an ellipse I11

= ab3 =4 I22 = a3b=4, we obtain


2 a2 = Mt b a+ 3 b3 :
(5.7.27)

Substitution for

into (5.7.17) and (5.7.19) yields

2 ; a2 u1 = ; X2X3 u2 = X1X3 u3 = b b2 + a2 X1X2


and

(5.7.28)

T11 = T22 = T33 = T12 = 0 2b2 X T = ;2a2 X : (5.7.29) T23 = 1 13 b2 + a2 b2 + a2 2 At any point (X1 X2 c) of a cross-section X3 = c n = (0 0 1) and the stress vector f is given
by

fi = Tij nj = b22+ a2 (;a2 X2 1i + b2X1 2i ):

f lies in the plane X3 = c, therefore, there is no normal stress at the point on the plane X3 = c. The magnitude of f gives the shear stress at the point.
Since

= 22 2 (a4X22 + b4X12 )1=2 : a +b


A point in the interior of the cross-section where takes on extremum values is given by

(5.7.30)

@ = @ = 0: @X1 @X2
120

This gives X1

= X2 = 0 and at this point = 0.

To nd points on the boundary where

may

have extremum values we rst write (5.7.30) as

ab2 . is maximum > b2 , it is obvious that is minimum at X2 = 0 and has the value 2 a2 + b2 2 2 a b . Thus, the maximum shear stress occurs at the extremities at X2 = b and has the value 2 a + b2
Since a2 of the minor axis of the ellipse, contrary to the intuitive expectation that the maximum shear stress would occur at points of maximum curvature. It is clear from (5.7.28)3 that the axial displacement of points in the rst and third quadrant will be along the the

= a22+ b2 (a2(a2 ; b2 )X22 + b4 a2 )1=2:

(5.7.31)

;X3 axis and that of points in the second and fourth quadrant will be along
X3-direction lie on a rectangular

X3-axis.

The points which have the same displacement in

hyperbola.

6
6.1 Constitutive Relation

The Linear Viscous Fluid

For a viscous material, the Cauchy stress tensor depends not only on the deformation gradient but also on its time derivative. That is

_ ): T = T(F F

(6.1.1)

Most uids are isotropic and homogeneous; an exception being a liquid crystal which is anisotropic. Here we will study isotropic uids only. For an isotropic uid, equation (6.1.1) reduces to

T = T( D)
and for a linear viscous uid, we have

(6.1.2)

T = ;p( )1 + (tr D)1 + 2 D:


Here p is the hydrostatic pressure, the bulk viscosity and

(6.1.3)

the shear viscosity. The word linear in

linear viscous uid signies that the viscous part of the stress (the last two terms on the right-hand
121

side of (6.1.3)) depends linearly upon the strain-rate tensor D. A perfect uid or an ideal gas has no viscosity, and therefore can be described by the constitutive relation

T = ;p( )1:
only on the present value of the deformation gradient. For an ideal gas,

(6.1.4)

Thus a perfect uid is a nonlinear elastic material in the sense that the Cauchy stress for it depends

p = RT
mations of an ideal gas at a constant temperature

(6.1.5)

where R is the universal gas constant and T is the temperature in degrees Kelvin. Thus for defor-

T = const. 1
and the value of the const. depends upon the gas and its temperature. Equation (6.1.3) for D = 0 gives

(6.1.6)

T = ;p( )1:

(6.1.7)

Hence in a uid at rest, the state of stress is a hydrostatic pressure. Note that the state of stress in a perfect uid or an ideal gas is always that of hydrostatic pressure whether or not it is being deformed. However, in a viscous uid, no shear stress exists if and only if it is at rest. This is sometimes taken as the denition of a viscous uid; viz. a viscous uid at rest can not support any shear stresses. On the other hand, a solid body when subjected to shear or tangential tractions can stay stationary. For a homogeneous uid the viscosities of the stress tensor T. and are constants. Equation (6.1.3) implies that the principal axes of the stretching tensor or the strain-rate tensor D coincide with the principal axes

Whereas the constitutive relation (5.2.1) or (5.2.2) for a linear elastic material describes well its innitesimal deformations, no such restriction is imposed on (6.1.3). Said differently, equation (6.1.3) describes deformations of a viscous uid for all values of the stretching tensor

D.

The

material characterized by equation (6.1.3) is called a Navier-Stokes uid. Usually uids and gases
122

are assumed to be incompressible. The constitutive relations for an incompressible ideal gas and an incompressible Navier-Stokes uid are respectively

T = ; p1 T = ; p1 + 2 D

(6.1.8) (6.1.9)

where the hydrostatic pressure p is not determined by the deformation of the uid. It is, however, determined by the boundary conditions. Even for a homogeneous uid, the pressure p is a function of the spatial coordinate x and time t. In (6.1.9) we have used the continuity condition, tr D = 0, for an incompressible uid. 6.2 Formulation of an Initial-Boundary-Value Problem One generally uses the spatial description of motion for a uid. Thus the balance of mass, and the balance of linear momentum for a compressible Navier-Stokes uid are

@ + @ ( vi ) = 0 @t @xi @vi + @vi v = ; @p @ + ( + ) @ 2 vk + @ 2 vi + g : i @t @xj j @ @xi @xi @xk @xk @xk

(6.2.1) (6.2.2)

Equation (6.2.2) is obtained by substituting into (4.1.14) for the acceleration from (3.6.3) and for the Cauchy stress from (6.1.3). Because of the presence of the convective part of the acceleration on the left-hand side of (6.2.2), equations governing the motion of a uid are nonlinear. In the referential description of motion used to describe the deformations of a solid, the acceleration is linear in displacements. Were we to use the referential description of motion for studying the deformations of a uid, the expression for the strain-rate tensor D in terms of the spatial gradients of the velocity eld will be more involved. The nonlinearity in (6.2.2) implies that we can no longer use the principle of superposition. Whereas for solids, the deformations caused by gravity are generally negligible and hence are ignored, gravity is the main driving force for a uid. A familiar example is the ow of a uid in a river. Note that equations governing the deformations of a solid were expressed in terms of displacements, those for a uid are written in terms of velocities. For an incompressible Navier-Stokes uid, equations expressing the balance of mass and the
123

balance of linear momentum are

The mass density

@vi = 0 (6.2.3) @xi @vi + @vi v = ; @p + @ 2 vi + g : (6.2.4) i @t @xj j @xi @xk @xk is a constant, and the pressure eld p is an arbitrary function of x and time t.

Equations (6.2.1) and (6.2.2) for a compressible Navier-Stokes uid or (6.2.3) and (6.2.4) for an incompressible Navier-Stokes uid are supplemented by the following initial and boundary conditions.

(x 0) = 0 (x) in

vi (x 0) = vi0 (x) in

(6.2.5)

;p ij + vk k ij + (vi j + vj i)]nj = fi (x t) on @1 vi(x t) = vip(x t) on @2


Here Equations (6.2.5) specifying the mass density and the velocity at time

(0 T ) (0 T ):
(6.2.6)

is the domain of study which may equal only a part of the region occupied by the uid.

t = 0 are the initial con-

ditions, and (6.2.6) prescribing surface tractions on a part of the boundary and velocities on the remainder of the boundary are boundary conditions. If the velocity eld is prescribed on the entire boundary, then for an incompressible uid, the pressure eld is indeterminate. For a steady ow, the eld variables time t. Thus not required. A ow is called irrotational if the spin tensor or the vorticity vanishes everywhere in the domain. Governing equations for the steady irrotational ow of an incompressible Navier-Stokes uid are

@ = 0 = @ v in equations (6.2.1) through (6.2.4), and initial conditions (6.2.5) are @t @t

p and v in the spatial description are independent of

@vi = 0 @xi @vi v = ; @p + g : @x j @x i


j i
does not inuence the steady irrotational ow of an incompressible Navier-Stokes uid.
124

(6.2.7) (6.2.8)

In writing (6.2.8) we have absorbed the constant mass density in p. Thus the viscosity of a uid

6.3 Examples
6.3.1 Steady Flow Between Two Parallel Plates

Consider the steady ow of a compressible Navier-Stokes uid between two parallel horizontal plates with the lower plate kept stationary and the upper one moved in the positive x1 -direction at a uniform speed of v0

m=s. Assume that the uid sticks to the plates and that vi (x t) = v(x2)

i1

where the x2 -axis is perpendicular to the plate surfaces, and the uid extends to innity in the

x3 -direction. Thus we have a two-dimensional problem in the x1 ; x2 plane with the uid owing only in the x1 -direction. It is therefore reasonable to assume that the mass density and the pressure eld also depend only upon x1 and x2 . For the presumed velocity eld, vi i = 0. Thus the motion
is isochoric and the mass density of a material particle does not change. Also, both the local and the convective parts of the acceleration identically vanish. Nontrivial equations governing the ow of the uid are

@ v=0 @x1 @p @ + @ 2 v 0= ;@ @x1 @x2 2 @p @ 0 = ; @ @x + g 2 v(x1 0) = 0 v(x1 h) = v0 Tij nj

(6.3.1) (6.3.2) (6.3.3) (6.3.4)

= p(x2 ) i1 Tij nj = ;p0 (x2 ) i1 (6.3.5) x1 =;L x1 =L Whereas equation (6.3.1) expresses the balance of mass, equations (6.3.2) and (6.3.3) are the reduced forms of the balance of linear momentum in the x1 and x2 directions. Equation (6.3.4) states the essential boundary conditions on the surfaces of the two plates, and equation (6.3.5) gives natural boundary conditions on the vertical surfaces x1 Equations (6.3.2) and (6.3.4) give

= L. It follows from (6.3.1) that = (x2 ).


(6.3.6)

2 v=x h v0 :
Thus the velocity eld is known. It follows from (6.3.3) that

d = g : dx2 @p=@
125

(6.3.7)

The determination of

as a function of x2 requires that the compressibility of the uid be known.

Said differently, the constitutive equation p = p(

) must be given. As an example, we take


(6.3.8)

p=c
where c is a constant. Then, equations (6.3.7) and (6.3.8) yield

= de c x2

(6.3.9)

where d is a constant. Equations (6.3.8) and (6.3.9) imply that p can depend only on x2 . The stress tensor in the uid computed from equations (6.1.3), (6.3.6), (6.3.8) and (6.3.9) is given by

0 Tij = ;cdegx2 =c ij + v (6.3.10) h ( 2j i1 + 2i j1) and is only a function of x2 . Thus p(x2 ) must equal p0 (x2 ) in equation (6.3.5); otherwise there
is no solution for the problem. That is, the pressure distribution must be same on every plane

x1 = const. From equations (6.3.5) and (6.3.10) we conclude that p(x2) = cdegx2=c :
That is, for the uid being studied, surface tractions on planes x1 (6.3.11)

= L can not be arbitrarily prescribed, but must be of the form (6.3.11). From the given value of p(x2 ), we nd the constant of integration d. The elds of density and velocity are respectively given by (6.3.9) and (6.3.6).
In linear elasticity, the uniquenes theorem guarantees that the solution obtained by a semiinverse method is the only solution of the problem. However, in uid mechanics, a uniqueness theorem can not be proved because of the nonlinear governing equations. It implies that there may be other solutions for the problem studied.
6.3.2 Steady Flow of an Incompressible Navier-Stokes Fluid down an Inclined Plane

In order to study this problem, it is more convenient to choose coordinate axes with x1 -axis along the inclined plane and x2 -axis perpendicular to it. We assume that the pressure and velocity elds are independent of x3 , and

vi(x1 x2 ) = v(x2 ) i1 :
126

(6.3.12)

That is, the uid is owing parallel to the inclined plane. Thus, the top surface of the uid will be parallel to the inclined plane. Such a ow is called laminar.

The balance of mass or the continuity equation, vi i

= 0, is identically satised.

Also, the

convective part of acceleration, vj vi j , vanishes. Thus equations expressing the balance of linear momentum are

@p + d2v + g sin 0 = ; @x dx2 1 2 @p 0 = ; @x ; g cos :


2
The associated boundary conditions are

(6.3.13) (6.3.14)

vi(x1 0) = 0 Tij nj Tij nj

x2 =h

= ;pa ni
(6.3.15)

= ti(x2 ) Tij nj = qi(x2 ): x1 =0 x1 =L That is, the uid adheres to the stationary inclined plane, and the surface tractions are prescribed
on the top surface and two arbitrarily chosen surfaces x1 A solution of equation (6.3.14) is

= 0 and x1 = L.
(6.3.16)

p = ; g cos x2 + f (x1 )
where f is an arbitrary smooth function of x1 . Substitution of (6.3.16) into (6.3.13) gives

df = d2v + g sin : dx1 dx2 2


127

(6.3.17)

Since the left-hand side of (6.3.17) is a function of x1 and the right-hand side a function of x2 , therefore each must equal a constant b. Thus

where c

e1 and e2 are constants of integration. Boundary condition (6.3.15)1 requires that e2 = 0. Tij = ; (; gx2 cos + bx1 + c) ij + ( i1 2j + j1 2i ) b ; g sin x2 + e1 :

f = bx1 + c 2 2 +e x +e v = b ; g sin x 2 1 2 2

(6.3.18)

Substitution from (6.3.12), (6.3.18) and (6.3.16) into the constitutive relation (6.1.9) gives

(6.3.19)

Equations (6.3.15)2 and (6.3.19) yield

;(; gh cos + bx1 + c) i2 +


Hence

i1

b ; g sin h + e = ;p : 1 a i2

(6.3.20)

; (; gh cos + bx1 + c) = ;pa b ; g sin h + e = 0


1
Because (6.3.21)1 must hold for all values of x1 , therefore

(6.3.21)

b = 0 c = gh cos + pa e1 = g sin h :
With these values of b

(6.3.22)

c and e1 , equation (6.3.19) gives


i1 2j + j 1 2i )

Tij = ;( g cos (;x2 + h) + pa ) ij + (


Noting that on the surface x1

g sin (;x + h) : 2

(6.3.23)

= 0 nj = ;

1j , equations (6.3.23) and (6.3.15) give 2i

( g cos (;x2 + h) + pa ) i1 +
One can similarly nd that qi (x2 )

g sin (x ; h) = t (x ): 2 i 2

(6.3.24)

= ;ti (x2 ). Unless normal and tangential tractions given by (6.3.24) are supplied on the planes x1 = const., the assumed ow vi = v (x2 ) i1 can not be
maintained on the inclined plane. The inclined plane is usually assumed to be innitely long and tractions necessary to maintain the ow are presumed to act on the planes x1
128

= const.

Das könnte Ihnen auch gefallen