You are on page 1of 28

Solid Mechanics: Linear Elasticity

N. H. Scott
(School of Mathematics, University of East Anglia,
Norwich, NR4 7TJ. United Kingdom)
1 Introduction
In a solid material (e.g. steel, rubber, wood, crystals, etc.) a deformation (i.e. a
change of shape) can be sustained only if a system of forces (loads) is applied to the
bounding surfaces, so setting up internally a distribution of stress (i.e. force per unit
area). The dening characteristic of an elastic material is that when these loads are
removed the body returns exactly to its original shape. The original state in which no
loads are applied is often called the natural state, or reference conguration.
Almost all engineering materials possess the property of elasticity provided the ex-
ternal loads are not too large. If the loads are increased beyond a certain limit, dierent
for each material, then the material will fail, either by fracture or by ow. In neither
case does the material return to its original shape when the loads are removed and we
say that the elastic limit has been exceeded. A solid material that fails by fracture is
said to be brittle and one that fails by ow is said to be plastic.
We do not consider atomic structure. We assume that matter is homogeneous and
continuously distributed over the material body, so that the smallest part of the body
possesses the same physical properties as the whole body. The theory of the mechanical
behaviour of such materials is called continuum mechanics. We consider therefore
only the macroscopic (large scale) behaviour of materials, which is adequate for most
engineering purposes, and ignore the microscopic (small scale) behaviour.
We further assume that the deformation is small so that the elastic material is lin-
early elastic. For almost all engineering materials the linear theory of elasticity holds
if the applied loads are small enough.
This unit discusses only the linear theory of elasticity.
2 The kinematics of deformation: strain
Kinematics is the study of motion (and deformation) without regard for the forces
causing it.
A material body B occupies a region B
0
of space at time t = 0, its reference cong-
uration, and a material particle has position vector X, with Cartesian components
X
i
, i = 1, 2, 3, with respect to the orthonormal vectors {e
1
, e
2
, e
3
}:
X =
3

i=1
X
i
e
i
.
1
Figure 1: Reference conguration Current conguration
The body undergoes a motion, or deformation, during which the material particle at
X when t = 0 is moved to
x = x(X, t) (2.1)
at time t and then has position vector
x =
3

i=1
x
i
e
i
,
with Cartesian components x
i
, i = 1, 2, 3. Using the Einstein summation conven-
tion, by which twice-occurring subscripts are summed over (from 1 to 3), we can write
X = X
i
e
i
, x = x
i
e
i
.
We write
x = X+u(X, t), x
i
= X
i
+ u
i
(X, t), i = 1, 2, 3, (2.2)
so that u(X, t) is the particle displacement with components u
i
, i = 1, 2, 3. The
displacement gradient, h, is the 3 3 matrix with components
h
ij
=
u
i
X
j
.
In the linear theory of elasticity we assume not only that the components u
i
are small
but also that each of the derivatives h
ij
is small. This has an important consequence:
since xX = u is small we can aord to replace coordinates X by x and write in place
of (2.2)
x = X+u(x, t), x
i
= X
i
+ u
i
(x, t), i = 1, 2, 3, (2.3)
and work in future only with the coordinates x
i
and the particle displacements u
i
(x, t).
2
The particle velocity is
v =
x
t

u
t

x
,
and the particle acceleration is similarly approximated by
a =

2
x
t
2


2
u
t
2

x
.
The deformation enters into the linear theory of elasticity only through the linear
strain tensor, e, which has components
e
ij
=
1
2
_
u
i
x
j
+
u
j
x
i
_
(2.4)
(i, j = 1, 2, 3). For example,
e
11
=
1
2
_
u
1
x
1
+
u
1
x
1
_
=
u
1
x
1
and e
12
=
1
2
_
u
1
x
2
+
u
2
x
1
_
.
A tensor is a linear map from one vector space to another; but all such maps can
be represented by matrices. So, for tensor, think matrix.
The strain tensor e is symmetric:
e
T
= e, e
ji
= e
ij
,
i.e. e
12
= e
21
, e
23
= e
32
, e
31
= e
13
. We see this from (2.4):
e
ji
=
1
2
_
u
j
x
i
+
u
i
x
j
_
=
1
2
_
u
i
x
j
+
u
j
x
i
_
= e
ij
.
The matrix (e
ij
) of strain components is therefore real and symmetric and so, by the
theory of linear algebra, has 3 real eigenvalues with corresponding unit eigenvectors
which are mutually orthogonal. The eigenvalues are known as the principal strains
and the unit eigenvectors are the principal axes of strain.
Triaxial stretch
Consider a unit cube aligned with the coordinate axes and subjected to the deformation
u
1
= e
1
x
1
, u
2
= e
2
x
2
, u
3
= e
3
x
3
, (2.5)
independent of time t, in which e
i
are the constant (small) strains. Note that e
i
> 0
corresponds to a stretch, and e
i
< 0 to a contraction.
From (2.5) we see that the strain components are given by
e
11
=
u
1
x
1
= e
1
, e
22
=
u
2
x
2
= e
2
, e
33
=
u
3
x
3
= e
3
,
with all the other components vanishing: e
12
= e
21
= e
23
= e
32
= e
31
= e
13
= 0. Thus
(e
ij
) =
_
_
_
e
1
0 0
0 e
2
0
0 0 e
3
_
_
_
3
Figure 2: Triaxial stretch
and we see that the principal strains are given by e
1
, e
2
, e
3
. Also, the principal axes of
strain are aligned with the coordinate axes of Figure 2 and have components
_
_
_
1
0
0
_
_
_,
_
_
_
0
1
0
_
_
_,
_
_
_
0
0
1
_
_
_,
respectively.
After the triaxial stretch the new volume is
(1 +e
1
)(1 +e
2
)(1 +e
3
) = 1 +e
1
+ e
2
+ e
3
+ e
1
e
2
+ e
2
e
3
+ e
3
e
1
+ e
1
e
2
e
3
.
Since the e
i
are small, the relative change in volume, known as the dilatation, is
= e
1
+ e
2
+ e
3
=
change in volume
original volume
.
But in this case e
1
= e
11
, e
2
= e
22
, e
3
= e
33
, so
= e
11
+ e
22
+ e
33
= tr e = e
kk
, (2.6)
where k is summed over. In fact, for any deformation, with e not necessarily diagonal,
the dilatation is given by (2.6). To see this we observe that any strain tensor e can be
diagonalized (because symmetric) by a suitable rotation of coordinates and then in the
new coordinates the deformation is necessarily of the form (2.5). So the dilatation is
given by (2.6). But tr e = e
kk
is an invariant quantity, unchanged by any rotation of
coordinates. Therefore (2.6) represents the dilatation for general e.
4
Conservation of mass
The mass m of the cube and the cuboid in Figure 2 are the same. Let
0
and be the
initial and nal densities:

0
=
m
1
, =
m
1 +
=
0
(1 + )
1
.
So, by the binomial theorem, with small, the conservation of mass reads
=
0
(1 ). (2.7)
If the volume increases ( > 0), the density decreases ( <
0
);
if the volume decreases ( < 0), the density increases ( >
0
).
Simple shear
Consider the same unit cube aligned with the coordinate axes but subjected to the shear
deformation
u
1
= x
2
, u
2
= 0, u
3
= 0, (2.8)
with a positive constant.
Figure 3: Simple shear
Then
u
1
x
2
= , with all other components of deformation gradient vanishing. So
e
12
=
1
2
_
u
1
x
2
+
u
2
x
1
_
=
1
2
( + 0) =

2
= e
21
,
with all other strain components vanishing:
(e
ij
) =
_
_
_
0
1
2
0
1
2
0 0
0 0 0
_
_
_.
For general strain tensor e the diagonal components e
11
, e
22
, e
33
are termed the
normal strains, whereas the o-diagonal elements e
12
= e
21
, e
23
= e
32
, e
31
= e
13
, are
termed the shear strains.
5
3 The theory of stress: equations of motion
The forces acting on a deformed body are of two kinds.
1. Body force, b, measured per unit mass, acts on volume elements, e.g. gravity,
inertial forces.
2. Contact force, t(n), measured per unit area.
Figure 4: Contact forces
Take an arbitrary surface element da with unit normal vector n. The material on
the side of da into which n points exerts a force t(n)da, across the surface element da,
on the material on the other side of da. The vector t(n) is the traction vector.
Example 1: Uniaxial tension A force per unit (deformed) area T is applied to a
Figure 5: Uniaxial tension
cuboid in the x
1
-direction causing extension in the x
1
-direction and (usually) lateral
contraction. Now
t(e
1
) = Te
1
, but t(e
2
) = 0,
so t(n) depends on n, even for xed x and t.
Figure 6: Simple shear
Example 2: Simple shear In this case t(e
2
) = Te
1
, so t(n) is not necessarily parallel
to n.
6
Consider a cuboid of deformed material with sides x
1
, x
2
, x
3
parallel to the coor-
dinate axes. The tractions on each face are as shown.
Figure 7: The stress components
The traction vector t(e
1
) on the face with normal e
1
has components
_
_
_

11

21

31
_
_
_,
the traction vector t(e
2
) on the face with normal e
2
has components
_
_
_

12

22

32
_
_
_,
the traction vector t(e
3
) on the face with normal e
3
has components
_
_
_

13

23

33
_
_
_.
These column vectors are put together to form the 3 3 matrix of components of the
stress tensor, :
(
ij
) =
_
_
_

11

12

13

21

22

23

31

32

33
_
_
_. (3.1)
The diagonal elements
11
,
22
,
33
are termed normal stresses and the o-diagonal
elements are termed shear stresses.
It can be shown that for arbitrary unit normal vector n = n
1
e
1
+n
2
e
2
+n
3
e
3
= n
i
e
i
,
the traction vector is given by
t(n) = n, t
i
=
ij
n
j
,
_
_
_
t
1
t
2
t
3
_
_
_ =
_
_
_

11

12

13

21

22

23

31

32

33
_
_
_
_
_
_
n
1
n
2
n
3
_
_
_, (3.2)
so the traction components t
i
are a linear combination of the unit surface normal com-
ponents n
i
.
7
The normal component of traction is
t
n
= t(n) n = t
i
n
i
=
ij
n
j
n
i
=
ij
n
i
n
j
, (3.3)
from (3.2). The normal stress t
n
is tensile when positive and compressive when
negative.
Figure 8: Normal and shear components of traction
The tangential component of traction is
t(n) t
n
n
with magnitude

n
= |t(n) t
n
n| = {|t(n)|
2
t
2
n
}
1
2
, (3.4)
called the shear stress.
Example 3 For the stress components at a point P
_
_
_
1 2 3
2 4 6
3 6 1
_
_
_
(a) nd the traction components at P on a plane whose outward unit normal has com-
ponents (
3
5
, 0,
4
5
),
(b) nd the normal and shear stresses at P on the given plane.
Solution (a) The traction components are
t(n) =
_
_
_
1 2 3
2 4 6
3 6 1
_
_
_
_
_
_
3
5
0
4
5
_
_
_ =
_
_
_
3
6
13
5
_
_
_
(b) From (3.3) the normal stress is
t
n
= n t(n) =
_
3
5
, 0,
4
5
_

_
_
_
3
6
13
5
_
_
_ =
97
25
.
From (3.4) the shear stress is

n
= |t(n)t
n
n| = {|t(n)|
2
t
2
n
}
1
2
=
_
3
2
+ 6
2
+
_
13
5
_
2

_
97
25
_
2
_1
2
=

22, 941
25
6.059.
8
The equations of motion
The balance of linear momentum
Figure 9: Balance of forces in the x
2
-direction
Consider an elementary cuboid of material with sides x
1
, x
2
, x
3
parallel to the
coordinate axes. The cuboid has volume v = x
1
x
2
x
3
. We calculate all the forces
acting on the element in the x
2
-direction. Unbalanced normal stress

22
x
2
x
2
acts on an
area x
1
x
3
, giving a normal force on the element in the x
2
-direction of

22
x
2
x
2
x
1
x
3
=

22
x
2
v. Unbalanced shear stresses

23
x
3
x
3
and

21
x
1
x
1
act on areas x
1
x
2
and
x
2
x
3
, respectively, giving shear forces in the x
2
-direction of

23
x
3
v and

21
x
1
v ,
respectively. The mass of the element is v and so the component of body force (per unit
mass) b
2
in the x
2
-direction contributes a force b
2
v on the element in that direction.
By Newtons second law the total of these forces can be equated to the rate of change
of linear momentum in the x
2
-direction:

21
x
1
v +

22
x
2
v +

23
x
3
v + b
2
v = v

2
u
2
t
2
.
On dividing by v we obtain the equation of motion

21
x
1
+

22
x
2
+

23
x
3
+ b
2
=

2
u
2
t
2
.
By similar arguments in the x
1
- and x
3
-directions we deduce a complete set of three
equations of motion:

11
x
1
+

12
x
2
+

13
x
3
+ b
1
=

2
u
1
t
2
,
9

21
x
1
+

22
x
2
+

23
x
3
+ b
2
=

2
u
2
t
2
, (3.5)

31
x
1
+

32
x
2
+

33
x
3
+ b
3
=

2
u
3
t
2
.
These equations are conveniently written in sux notation as

ij
x
j
+ b
i
=

2
u
i
t
2
(3.6)
for each i = 1, 2, 3, and summing over j.
In the static case, in which no quantities vary with time t then (3.6) reduce to the
equilibrium equations

ij
x
j
+ b
i
= 0. (3.7)
The balance of angular momentum
Figure 10: Shear stresses
Taking moments about the origin O shows that

21
=
12
.
Similarly, we obtain

31
=
13
,
23
=
32
.
So the balance of angular momentum reduces to the
symmetry of the stress tensor:

T
= ,
ji
=
ij
. (3.8)
(Recall also the symmetry of the strain tensor e).
Properties of the stress tensor,
It follows from this symmetry that has three real eigenvalues with corresponding real
unit eigenvectors. These three eigenvalues are termed the principal stresses and the
corresponding unit eigenvectors are termed the principal axes of stress. They are
mutually orthogonal.
10
If the stress has diagonal form
_
_
_

1
0 0
0
2
0
0 0
3
_
_
_
then the principal stresses are
1
,
2
,
3
with corresponding principal axes of stress
e
1
, e
2
, e
3
, along the coordinate axes. The traction on a plane with normal components
(n
1
, n
2
, n
3
) is
t(n) =
_
_
_

1
0 0
0
2
0
0 0
3
_
_
_
_
_
_
n
1
n
2
n
3
_
_
_ =
_
_
_

1
n
1

2
n
2

3
n
3
_
_
_.
The normal stress is therefore, from (3.3),
t
n
=
1
n
2
1
+
2
n
2
2
+
3
n
2
3
(3.9)
and the squared shear stress can, from (3.4), be shown to be given by

2
n
= (
1

2
)
2
n
2
1
n
2
2
+ (
1

3
)
2
n
2
1
n
2
3
+ (
2

3
)
2
n
2
2
n
2
3
. (3.10)
If the stress takes the form
= pI,
ij
= p
ij
in which
ij
denotes the Kronecker delta

ij
=
_
1 i = j,
0 i = j,
in other words the components of the unit tensor I, it is said to be spherical (or
hydrostatic) and p is the pressure. For a surface segment with unit normal n
t(n) = n = pIn = pn, t
i
= pn
i
,
so that the normal stress and shear stress are, from (3.3) and (3.4),
t
n
= (pn) n = pn n = p and
n
= 0.
If the stress is spherical, therefore, the traction is purely normal on every surface element,
so that there is no shear stress on any surface element.
Example 4 For the matrix of stress components of Example 3, namely,
(
ij
) =
_
_
_
1 2 3
2 4 6
3 6 1
_
_
_,
nd the principal stresses and the corresponding principal axes of stress.
11
Solution Let denote the eigenvalues and n the eigenvectors of the stress tensor .
Then, by denition, n = n, so that ( I)n = 0, or, in components,
_
_
_

11

12

13

21

22

23

31

32

33

_
_
_
_
_
_
n
1
n
2
n
3
_
_
_ =
_
_
_
0
0
0
_
_
_.
There are non-trivial solutions (i.e. n = 0) if and only if the determinant of coecients
vanishes:

11

12

13

21

22

23

31

32

33

= 0.
In the present example this condition becomes

1 2 3
2 4 6
3 6 1

= 0.
Expand by rst row:
(1 ){(4 )(1 ) 36} 2{2(1 ) 18} + 3{12 3(4 )} = 0.
Simplify the curly brackets:
(1 )(
2
5 32) 2(2 16) + 9 = 0.
Multiply out and collect terms:

3
+ 6
2
+ 40 = 0.
Change overall sign and observe that = 0 is a root:
(
2
6 40) = 0, and factorise: ( 10)( + 4) = 0.
Therefore the principal stresses are

1
= 0,
2
= 10,
3
= 4.
We must now nd the corresponding unit eigenvectors.
For =
1
= 0 we must nd components n
i
such that
_
_
_
1 2 3
2 4 6
3 6 1
_
_
_
_
_
_
n
1
n
2
n
3
_
_
_ =
_
_
_
0
0
0
_
_
_.
2nd equation same as 1st. 1st and 3rd give n
3
= 0, and then n
1
= 2, n
2
= 1 are
possible solutions. The unit vector n
1
therefore has components 5

1
2
(2, 1, 0).
For =
2
= 10 we nd the unit vector n
2
has components 70

1
2
(3, 6, 5).
For =
3
= 4 we nd the unit vector n
3
has components 14

1
2
(1, 2, 3).
12
4 Linear isotropic elasticity: constitutive equations
and homogeneous deformations
The constitutive equation of isotropic linear elasticity (the stress-strain law) is

ij
= e
pp

ij
+ 2e
ij
. (4.1)
known as the generalized Hookes law, with e
ij
dened by (2.4). The material con-
stants (or elastic moduli) and are known as the Lame moduli and must be
measured experimentally for each material.
In the absence of body force the equations of equilibrium (3.7) are

ij
x
j
= 0. (4.2)
Homogeneous deformations
A deformation is said to be homogeneous if the strain tensor e is independent of x, i.e.
spatially uniform. From (4.1) the stress tensor is also uniform in space. Thus (4.2)
is satised trivially. Therefore, any homogeneous deformation satises the equations of
equilibrium and so is possible in an elastic material, in the absence of body force.
We examine various examples of homogeneous deformation.
(a) Uniform dilatation
An isotropic elastic sphere of radius a centred on the origin undergoes the uniform
dilatation
u
1
= x
1
, u
2
= x
2
, u
3
= x
3
, (or u = x), (4.3)
where is a dimensionless constant; > 0 for expansion and < 0 for contraction.
Now
_
u
i
x
j
_
=
_
_
_
0 0
0 0
0 0
_
_
_
and so
e
ij
=
ij
.
The dilatation is = e
pp
= 3, and from (4.1),

ij
= 3
ij
+ 2
ij
= (3 + 2)
ij
so
11
=
22
=
33
= p, where p = (3 +2), with all shear stresses vanishing.
We have
p = K
where
K :=
1
3

pp
e
pp
, or K =
p

= +
2
3
(4.4)
is the bulk modulus of the material.
13
(b) Simple shear
The displacements are
u
1
= x
2
, u
2
= 0, u
3
= 0,
so that
(e
ij
) =
_
_
_
0
1
2
0
1
2
0 0
0 0 0
_
_
_ and e
pp
= 0.
So from (4.1),
Figure 11: Simple shear

ij
= 2e
ij
=
_
_
_
0 0
0 0
0 0 0
_
_
_.
Therefore,
12
= , e
12
=
1
2
are the only non-zero stresses and strains.
The quantity
=

12
2e
12
is known as the shear modulus of the material.
(c) Simple extension (uniaxial tension)
Suppose that a cylindrical rod lies with its generators parallel to the x
1
-direction and
suers a uniform tension T at each end:
(
ij
) =
_
_
_
T 0 0
0 0 0
0 0 0
_
_
_.
We assume a displacement
u
1
= x
1
, u
2
= x
2
, u
3
= x
3
,
where > 0 corresponds to extension of the rod and > 0 to lateral contraction. So
(e
ij
) =
_
_
_
0 0
0 0
0 0
_
_
_, e
pp
= 2.
14
Figure 12: Uniaxial tension
From (4.1)

11
= T = ( 2) + 2,
22
=
33
= 0 = ( 2) 2.
From the second equation we get and then from the rst we get T:
=

2( + )
, T =
(3 + 2)
+
.
We dene the Youngs modulus E by
E :=

11
e
11
, so E =
T

=
(3 + 2)
+
(4.5)
and Poissons ratio by
:=
e
22
e
11
, so =

=

2( + )
, (4.6)
both in terms of and . E is the tension per unit axial extension and is the transverse
contraction per unit axial extension.
Elastic constants
We have introduced elastic constants K, E, in addition to the Lame constants ,
and have provided clear mechanical interpretations of K, E, and . Any three of
, , K, E, can be expressed in terms of the other two by manipulating (4.4), (4.5)
and (4.6).
We shall suppose, on physical grounds, that
K > 0, > 0 : (4.7)
K > 0 implies that a compressive pressure produces a decrease in volume
> 0 implies a shear stress produces shear strain in the same direction,
not in the opposite direction.
By manipulating (4.4)(4.6) we obtain
E =
9K
3K +
, =
3K 2
2(3K + )
(4.8)
15
and so inequalities (4.7) imply that
E > 0, 1 < <
1
2
. (4.9)
To obtain (4.9)
2
, write (4.8)
2
as
=
3
K

2
2
_
3
K

+ 1
_
(4.10)
and observe that 1 as
K

0, and
1
2
as
K

.
The inequalities (4.7) imply that
(a) pressure produces a decrease in volume in dilatation (K > 0)
(b) the shear is in the same direction as the shear stress
in simple shear ( > 0)
(c) axial tension results in axial elongation in simple extension (E > 0)
Thus inequalities (4.7) ensure physically reasonable response of the material in the three
homogeneous deformations discussed above. However, (4.9)
2
permits the possibility of
axial extension being accompanied by lateral expansion ( < 0). But no known isotropic
material responds to simple extension in this way. Therefore in practice 0 < <
1
2
.
The strain-stress law. Deviatoric tensors
The stress-strain law (4.1) is repeated here for convenience:

ij
= e
pp

ij
+ 2e
ij
.
Taking the trace of both sides gives

pp
= (3 + 2)e
pp
. (4.11)
Thus
ij
=

3 + 2

kk

ij
+ 2e
ij
and so
e
ij
=
1
2
_


3 + 2

pp

ij
+
ij
_
(4.12)
which is the strain-stress law, inverse to (4.1).
The deviatoric stress and deviatoric strain are dened by

ij
=
ij

1
3

pp

ij
, e

ij
= e
ij

1
3
e
pp

ij
, (4.13)
respectively, and have the property that their traces vanish:

pp
= 0, e

pp
= 0. (4.14)
On substituting for
ij
and e
ij
from (4.13) into (4.1) we nd that the stress-strain law
(4.1) may be written

pp
= 3Ke
pp
,

ij
= 2e

ij
, (4.15)
involving the bulk modulus K and the shear modulus .
16
From (4.15)
1
we may recover our previous denition (4.4) of bulk modulus:
K =
1
3

pp
e
pp
.
Equation (4.15)
2
makes clear the proportionality of shear stresses and strains, connected
by the shear modulus .
Incompressibility
An incompressible material is one whose volume cannot be changed, though it can be
changed in shape, i.e. distorted.
In an incompressible material, therefore, every deformation is such that the dilatation
vanishes:
= e
pp
= 0.
The uniform dilatation (4.3), i.e. u = x, cannot occur in an incompressible material
as e
pp
= 0 implies = 0.
For arbitrary hydrostatic pressure p there is no change of volume and so the stress-
strain law (4.1) is replaced by

ij
= p
ij
+ 2e
ij
(4.16)
in which p(x, t) is an arbitrary pressure, not dependent on the strains e
ij
.
The bulk modulus K is dened at (4.4) by
K =
p
e
pp
= +
2
3
.
In the limit of incompressibility e
pp
0 and so
K , (4.17)
and is unaltered.
For the simple shear discussed previously, e
pp
= 0 and so simple shear is possible in
any incompressible material.
For the uniaxial tension considered before:
u
1
= x
1
, u
2
= x
2
, u
3
= x
3
to be possible in an incompressible material the dilatation must vanish, so that
e
pp
= 2 = 0

=
1
2
.
But Poissons ratio is given by (4.6) to be
=
e
22
e
11
=

=
1
2
,
so that
=
1
2
(4.18)
for all incompressible materials. We can see this another way from (4.6)
=

2( + )
and for incompressibility , so that
1
2
, as at (4.18).
17
5 Conservation of energy: the strain energy func-
tion
The equations of motion (3.6) are

ij
x
j
+ b
i
=

2
u
i
t
2
. (5.1)
Multiply by the velocity v
i
= u
i
/t and sum over i:

ij
x
j
v
i
+ b
i
v
i
=
v
i
t
v
i
.
Thus

x
j
(
ij
v
i
)
ij
v
i
x
j
+ b
i
v
i
=
1
2


t
(v
i
v
i
). (5.2)
Now
v
i
x
j
=

x
j
u
i
t
=

t
u
i
x
j
=
e
ij
t
+

ij
t
,
where the strain tensor e and rotation are dened by
e
ij
=
1
2
_
u
i
x
j
+
u
j
x
i
_
,
ij
=
1
2
_
u
i
x
j

u
j
x
i
_
.
Then

ij
v
i
x
j
=
ij
_
e
ij
t
+

ij
t
_
,
But is skew-symmetric, i.e.
T
= or
ji
=
ij
, so

t
also is skew-symmetric
and it follows that
ij

ij
t
= 0.
Then (5.2) may be written

x
j
(
ij
v
i
) + b
i
v
i
=
ij
e
ij
t
+
1
2


t
(v
2
).
Integrate over an arbitrary sub-region R
t
of the body B
t
and use the divergence theorem:
_
Rt

ij
v
i
n
j
da +
_
Rt
b v dv =
_
Rt

ij
e
ij
t
dv +
_
Rt
1
2


t
(v
2
) dv.
But t(n) = n and is assumed constant to this order:
_
Rt
t(n) v da +
_
Rt
b v dv =
_
Rt

ij
e
ij
t
dv +

t
_
Rt
1
2
v
2
dv. (5.3)
The terms of this equation are interpreted as follows:
1st term: rate of working of surface tractions
2nd term: rate of working of body forces
3rd term: stress power
4th term: rate of change of kinetic energy
18
The stress power per unit volume is dened by
P :=
ij
e
ij
t
. (5.4)
Suppose this is wholely derived as the rate of change of a single function W(e), the
strain energy, measured per unit volume, which depends on the strain e:
W
t
=
ij
e
ij
t
.
Now by the chain rule
W
e
ij
e
ij
t
=
ij
e
ij
t
so that
_
W
e
ij

ij
_
e
ij
t
= 0.
But at each position x and corresponding value of e,
e
ij
t
may be selected arbitrarily,
so since the brackets do not depend on
e
ij
t
, we may conclude:

ij
=
W
e
ij
. (5.5)
Thus W(e) is a potential function for the stress . It is often termed the potential
energy. These ideas carry over into nonlinear elasticity.
The strain energy in isotropic linear elasticity
From (4.1) we see that the stress components
ij
are a linear combination of the strain
components e
ij
. So from (5.5), W(e) must be a homogeneous quadratic in the strain
components.
From Eulers result on the partial dierentiation of a homogeneous function of de-
gree n:
[Namely, if (x
i
) =
n
(x
i
) then
M

i=1
x
i

x
i
= n]
W
e
ij
e
ij
= 2W
so from (5.5)
W =
1
2

ij
e
ij
. (5.6)
Using the generalized Hookes law (4.1) allows the strain energy to be expressed entirely
in terms of the strain components e
ij
:
W =
1
2
(e
pp
)
2
+ e
ij
e
ij
. (5.7)
19
Written out in full
W =
1
2
(e
11
+ e
22
+ e
33
)
2
+ {e
2
11
+ e
2
22
+ e
2
33
+ 2e
2
12
+ 2e
2
23
+ 2e
2
31
}. (5.8)
On physical grounds we expect the quadratic form (5.7) to be positive denite:
W(e) 0, with W(e) = 0 only if e = 0. (5.9)
Any straining of the body (e = 0) requires work to be done on the body (W > 0):
W < 0 for some e = 0 would imply that starting from an unstressed state of rest work
could spontaneously be done on the surroundings.
We now seek necessary and sucient conditions on and for W(e) given by (5.8)
to be positive denite. From (4.4) the bulk modulus is K = +
2
3
and elimination of
from (5.8) in favour of K gives, after much manipulation,
W =
1
2
K(e
11
+ e
22
+ e
33
)
2
+
1
3

_
(e
11
e
22
)
2
+ (e
22
e
33
)
2
+ (e
33
e
11
)
2
_
+ 2
_
e
2
12
+ e
2
23
+ e
2
31
_
. (5.10)
For the uniform dilatation e
11
= e
22
= e
33
= , e
ij
= 0 for i = j, (5.10) becomes
W =
9K
2

2
,
so that W > 0 for = 0 requires K > 0.
For the simple shear e
12
= e
21
= /2, all other e
ij
vanishing, (5.10) becomes
W =

2

2
,
so that W > 0 for = 0 requires > 0. Therefore, necessary conditions for the
positive deniteness of W are
K > 0, > 0. (5.11)
These are also sucient as they ensure that, from (5.10), W is expressed as a sum of
squares with positive coecients such that W(e) = 0 requires e = 0.
But conditions (5.11) are precisely those adopted before, on dierent physical grounds,
see (4.7) and the following text.
At (4.13)
2
we introduced the strain deviator
e

ij
= e
ij

1
3
e
pp

ij
e

pp
= 0, (5.12)
in terms of which it may be shown that the strain energy W, taken in the form (5.7),
may be written
W =
1
2
K(e
pp
)
2
+ e

ij
e

ij
. (5.13)
Conditions (5.11) may also be deduced from this form of W.
20
6 The torsion problem
We consider an isotropic elastic cylinder of length l and arbitrary cross-section, placed
with its generators parallel to the x
3
-axis and with the origin 0 in one end face, as shown
in Figure 12. The cylinder is twisted about 0x
3
by an amount per unit length by the
application of shear tractions to its end faces, the curved surface remaining stress free.
The total angle of twist, l, is suciently small for its square to be neglected.
Figure 12: Cylinder under torsion Figure 13: The cross-section S
Figure 13 gives a plan view of the typical right cross-section S indicated in Figure 12.
If P is a typical point of S and 0

the point of intersection of 0x


3
with S, the deformation
rotates the line 0

P anticlockwise through an angle x


3
, as shown. The components
of the displacement of P in the x
1
- and x
2
-directions are therefore, with r = 0

P,
x
1
= r cos and x
2
= r sin ,
u
1
= r cos( + x
3
) r cos = r cos cos x
3
r sin sin x
3
r cos
= r sin (x
3
) + O(x
3
)
2
= x
2
x
3
,
u
2
= r sin( + x
3
) r sin = r sin cos x
3
+ r cos sin x
3
r sin
= r cos (x
3
) + O(x
3
)
2
= x
1
x
3
.
_

_
(6.1)
It follows from (6.1) and (2.4) that e
11
= e
22
= 0. We assume that e
33
= 0 also,
the torsional deformation consisting entirely of shear with no axial extension. Thus
u
3
/x
3
= 0, implying that u
3
is a function of x
1
and x
2
only:
u
3
= (x
1
, x
2
).
The function , specifying the axial deviation of S from its initially plane form, is
called the warping function. The displacements of the simple torsion deformation are
21
therefore
u
1
= x
2
x
3
, u
2
= x
1
x
3
, u
3
= (x
1
, x
2
), (6.2)
with strain components
(e
ij
) =

2
_
_
_
_
_
_
_
_
_
_
0 0

x
1
x
2
0 0

x
2
+ x
1

x
1
x
2

x
2
+ x
1
0
_
_
_
_
_
_
_
_
_
_
satisfying e
pp
= 0, so that from (4.1) the stress components are
(
ij
) =
_
_
_
_
_
_
_
_
_
_
0 0

x
1
x
2
0 0

x
2
+ x
1

x
1
x
2

x
2
+ x
1
0
_
_
_
_
_
_
_
_
_
_
. (6.3)
The equilibrium equations (3.7) in the absence of body force are

ij
x
j
= 0. (6.4)
The i = 1 and i = 2 equations are satised trivially and the i = 3 equation gives

x
1
_

x
1
x
2
_
+

x
2
_

x
2
+ x
1
_
=

2

x
2
1
+

2

x
2
2
=
2
= 0, (6.5)
the warping function thus being a harmonic function.
The outward unit normal n to C, the boundary of S, has components (n
1
, n
2
, 0),
see Figure 13. The components of the stress vector acting on the curved surface of the
cylinder are, from (3.2) and (6.3),
t(n) = n or
_
_
_
0 0
31
0 0
32

31

32
0
_
_
_
_
_
_
n
1
n
2
0
_
_
_ =
_
_
_
0
0

31
n
1
+
32
n
2
_
_
_.
Since this surface is stress free,

31
n
1
+
32
n
2
= 0 on C (6.6)
and substituting for the stress components from (6.3) gives

x
1
n
1
+

x
2
n
2
= x
2
n
1
x
1
n
2
on C,
or

n
= x
2
n
1
x
1
n
2
on C, (6.7)
/n being the directional derivative of along the outward normal to C. The warping
function thus satises the Neumann problem consisting of the plane Laplace equation
(6.5) in S together with the boundary condition (6.7) on C.
22
The twisting torque
The outward unit normal to the end face x
3
= l has components (0, 0, 1). The traction
vector acting on this face therefore has components
i3
and the components of the
resultant force F acting on the face are,
F
i
=
_ _
S

i3

x
3
=l
da.
From (6.3),
13
and
23
are independent of x
3
and
33
= 0 so
F
1
=
_ _
S

13
da, F
2
=
_ _
S

23
da, F
3
= 0.
From the i = 3 equilibrium equation (6.4) and the stress (6.3), and its symmetry

31
x
1
+

32
x
2
=

13
x
1
+

23
x
2
= 0.
It follows that
(x
1

13
)
x
1
+
(x
1

23
)
x
2
= x
1
_

13
x
1
+

23
x
2
_
+
13
=
13
,
(x
2

13
)
x
1
+
(x
2

23
)
x
2
= x
2
_

13
x
1
+

23
x
2
_
+
23
=
23
,
and
F
1
=
_ _
S
_
(x
1

13
)
x
1
+
(x
1

23
)
x
2
_
da =
_
C
x
1
(
31
n
1
+
32
n
2
) ds = 0,
F
2
=
_ _
S
_
(x
2

13
)
x
1
+
(x
2

23
)
x
2
_
da =
_
C
x
2
(
31
n
1
+
32
n
2
) ds = 0,
using the two-dimensional divergence theorem and the boundary condition (6.6). Thus
F = 0,
and, in exactly the same way, the resultant force on the end face x
3
= 0 is zero.
The torque per unit area about 0 acting on the face x
3
= l is
x t(e
3
) =

e
1
e
2
e
3
x
1
x
2
l

13

23

33

,
and so, remembering that
33
= 0, the resultant torque M acting on the face x
3
= l has
components
M
1
=
__
S
(x
2

33
l
32
)|
x
3
=l
da = lF
2
= 0,
M
2
=
__
S
(l
31
x
1

33
)|
x
3
=l
da = lF
1
= 0,
M
3
=
__
S
(x
1

32
x
2

31
)|
x
3
=l
da
=
_ _
S
_
x
1
_

x
2
+ x
1
_
x
2
_

x
1
x
2
__
da = D,
23
where
D =
_ _
S
_
x
2
1
+ x
2
2
+ x
1

x
2
x
2

x
1
_
da. (6.8)
Thus M = De
3
, where e
3
is the unit vector in the direction of x
3
increasing, and the
resultant torque on the end face x
3
= 0 is similarly De
3
.
The cylinder is therefore maintained in equilibrium by equal and opposite twisting
torques of magnitude D acting about 0x
3
on the end faces. The quantity D, dened
by (6.8), is called the torsional rigidity of S. D/ has physical dimension (length)
4
and depends only on the form of the cross-section S.
The Prandtl stress function
Since is a plane harmonic function, there exists an analytic function f of the complex
variable z = x
1
+ ix
2
such that
(x
1
, x
2
) = e f(z).
Let
(x
1
, x
2
) = mf(z),
so that
f(x
1
+ ix
2
) = (x
1
, x
2
) + i(x
1
, x
2
).
Then and are conjugate harmonic functions satisfying the Cauchy-Riemann equa-
tions

x
1
=

x
2
,

x
2
=

x
1
. (6.9)
The unit vector t tangential to C in the anticlockwise sense has components
t
1
= n
2
, t
2
= n
1
, t
3
= 0, (6.10)
see Figure 13. Using (6.9) and (6.10) we can rewrite (6.7) as follows:

x
1
n
1
+

x
2
n
2
= x
2
n
1
x
1
n
2
on C,

x
2
t
2
+
_

x
1
_
(t
1
) = x
2
t
2
x
1
(t
1
) on C,

x
1
t
1
+

x
2
t
2
= x
1
t
1
+ x
2
t
2
on C.
Now by denition

x
1
t
1
+

x
2
t
2
= t =

s
,
where s is arc length along C in the positive (anti-clockwise) sense. So (6.7) nally
reduces to

s
=
1
2

s
(x
2
1
+ x
2
2
) on C.
The Neumann problem (6.5), (6.7) is thus equivalent to the Dirichlet problem

2
= 0 in S,
=
1
2
(x
2
1
+ x
2
2
) + constant on C.
_
_
_
(6.11)
24
Simpler results are obtained by introducing the Prandtl stress function
=
1
2
(x
2
1
+ x
2
2
). (6.12)
From (6.11), satises the Poisson equation

2
= 2 in S, (6.13)
and the boundary condition
=
0
on C, (6.14)
where
0
is a constant.
In terms of , the non-zero shear stresses are given from (6.3), (6.9) and (6.12) by

23
=
_

x
2
+ x
1
_
=
_

x
1
x
1
_
=

x
1

31
=
_

x
1
x
2
_
=
_

x
2
x
2
_
=

x
2
,
_

_
(6.15)
and the torsional rigidity from (6.8), (6.9) and (6.12) by
D =
_ _
S
_
x
2
1
+ x
2
2
x
1

x
1
x
2

x
2
_
da =
_ _
S
_
x
1

x
1
+ x
2

x
2
_
da. (6.16)
Since
x
1

x
1
+ x
2

x
2
=
(x
1
)
x
1
+
(x
2
)
x
2
2,
the application of the two-dimensional divergence theorem gives
_ _
S
_
x
1

x
1
+ x
2

x
2
_
da = 2
_ _
S
da +
_
C
(x
1
n
1
+ x
2
n
2
)ds
= 2
_ _
S
da +
_
C

0
(x
1
n
1
+ x
2
n
2
)ds,
where (6.14) has been used. Thus
D = 2
_ _
S
da
_
C

0
(x
1
n
1
+ x
2
n
2
)ds. (6.17)
If S is simply connected, so that C is a single closed curve we can replace by
0
.
Equations (6.13), (6.15) and (6.16) are unchanged and (6.14) and (6.17) simplify to
= 0 on C, (6.18)
D = 2
_ _
S
da. (6.19)
25
Figure 14: The elliptical cylinder
Examples
1. Torsion of an elliptical cylinder
Suppose that C is an ellipse and that 0x
3
is the line of centres of the right cross-sections.
If a
1
and a
2
are the semi-axes of the ellipse in the x
1
- and x
2
-directions, the equation of
C is
x
2
1
a
2
1
+
x
2
2
a
2
2
= 1.
The function
(x
1
, x
2
) = A
_
x
2
1
a
2
1
+
x
2
2
a
2
2
1
_
clearly satises the boundary condition (6.18), i.e. vanishes on C, and

2
=

2

x
2
1
+

2

x
2
2
= A
_
2
a
2
1
+
2
a
2
2
_
.
The Poisson equation (6.13) therefore holds if
A
_
2
a
2
1
+
2
a
2
2
_
= 2 A =
a
2
1
a
2
2
a
2
1
+ a
2
2
,
and so the Prandtl stress function for the elliptical cylinder is
=
a
2
1
a
2
2
a
2
1
+ a
2
2
_
x
2
1
a
2
1
+
x
2
2
a
2
2
1
_
=
a
2
1
a
2
2
a
2
2
x
2
1
a
2
1
x
2
2
a
2
1
+ a
2
2
.
From (6.12), i.e. = +
1
2
(x
2
1
+ x
2
2
),
(x
1
, x
2
) =
a
2
1
a
2
2
a
2
2
x
2
1
a
2
1
x
2
2
a
2
1
+ a
2
2
+
1
2
(x
2
1
+ x
2
2
)
=
1
2(a
2
1
+ a
2
2
)
(2a
2
1
a
2
2
2a
2
2
x
2
1
2a
2
1
x
2
2
+ a
2
1
x
2
1
+ a
2
1
x
2
2
+ a
2
2
x
2
1
+ a
2
2
x
2
2
)
=
1
2(a
2
1
+ a
2
2
)
_
(a
2
1
a
2
2
)(x
2
1
x
2
2
) + 2a
2
1
a
2
2
_
= m
1
2(a
2
1
+ a
2
2
)
_
(a
2
1
a
2
2
)iz
2
+ 2ia
2
1
a
2
2
_
26
where z = x
1
+ ix
2
and so z
2
= x
2
1
x
2
2
+ 2ix
1
x
2
and iz
2
= i(x
2
1
x
2
2
) 2x
1
x
2
. The
warping function is therefore
(x
1
, x
2
) = e
1
2(a
2
1
+ a
2
2
)
_
(a
2
1
a
2
2
)iz
2
+ 2ia
2
1
a
2
2
_
=
a
2
1
a
2
2
a
2
1
+ a
2
2
x
1
x
2
.
When a
1
> a
2
the warping pattern is as shown in Figure 15. The curves = constant
are rectangular hyperbolae.
Figure 15: The warping pattern for an elliptical cylinder
The torsional rigidity D is, from (6.19),
D = 2
_ _
S
da =
2a
2
1
a
2
2
a
2
1
+ a
2
2
_ _
S
_
1
x
2
1
a
2
1

x
2
2
a
2
2
_
dx
1
dx
2
.
Calculation of D. Because S is elliptical we transform the double integral using
modied polar coordinates
x
1
= a
1
r cos , x
2
= a
2
r sin .
Since
x
2
1
a
2
1
+
x
2
2
a
2
2
= r
2
the limits of integration are 0 r 1, 0 2.
The Jacobian is
(x
1
, x
2
)
(r, )
=

x
1
r
x
1

x
2
r
x
2

a
1
cos a
1
r sin
a
2
sin a
2
r cos

= a
1
a
2
r(cos
2
+ sin
2
) = a
1
a
2
r,
so that dx
1
dx
2
a
1
a
2
rdrd. Then the double integral is
_ _
S
_
1
x
2
1
a
2
1

x
2
2
a
2
2
_
dx
1
dx
2
=
_
2
=0
_
1
r=0
_
1
a
2
1
r
2
cos
2

a
2
1

a
2
2
r
2
sin
2

a
2
2
_
a
1
a
2
rdrd
27
= a
1
a
2
_
2
=0
_
1
r=0
_
1 r
2
_
rdrd = 2a
1
a
2
_
r
2
2

r
4
4
_
1
0
=
a
1
a
2
2
.
Therefore
D =
2a
2
1
a
2
2
a
2
1
+ a
2
2

a
1
a
2
2
=
a
3
1
a
3
2
a
2
1
+ a
2
2
,
the torsional rigidity of an elliptical cylinder.
Torsion of a circular cylinder In the case of a circular cylinder, a
1
= a
2
= a and
(x
1
, x
2
) =
1
2
(a
2
x
2
1
x
2
2
),
(x
1
, x
2
) = 0,
D =
1
2
a
4
.
Because = 0, there is no warping when the circular cylinder is twisted about its central
axis.
2. Torsion of a circular tube
Figure 16: The circular tube
We consider the cross-section S bounded by the concentric circles C
1
, C
2
of radii a
and b (< a). The axis of torsion is once again the axis of symmetry.
The function
(x
1
, x
2
) =
1
2
(a
2
x
2
1
x
2
2
),
obtained above for the circular cylinder, satises the Poisson equation (6.13), vanishes
on C
1
and is constant on C
2
. It is therefore an acceptable Prandtl stress function for
the circular tube and, as before, = 0, so that the right cross-sections are unwarped.
Because the boundary of the tube consists of disjoint curves we cannot use the fomula
(6.19) for the torsional rigidity. Instead, we obtain from (6.16) the expression
D =
_ _
S
_
x
2
1
x
2
2
_
da =
_
a
b
r
2
.2rdr = 2
_
1
4
r
4
_
a
b
=
1
2
(a
4
b
4
)
for the torsional rigidity of a circular tube.
28

You might also like