Sie sind auf Seite 1von 116

1 Tensors

1.1 Introduction
As seen previously in the introductory chapter, the goal of continuum mechanics is to
establish a set of equations that governs a physical problem from a macroscopic
perspective. The physical variables featuring in a problem are represented by tensor fields,
in other words, physical phenomena can be shown mathematically by means of tensors
whereas tensor fields indicate how tensor values vary in space and time. In these equations
one main condition for these physical quantities is they must be independent of the
reference system, i.e. they must be the same for different observers. However, for matters
of convenience, when solving problems, we need to express the tensor in a given
coordinate system, hence we have the concept of tensor components, but while tensors are
independent of the coordinate system, their components are not and change as the system
change.
In this chapter we will learn the language of TENSORS to help us interpret physical phenomena.
These tensors can be classified according to the following order:
Zeroth-Order Tensors (Scalars): Among some of the quantities that have magnitude but
not direction are e.g.: mass density, temperature, and pressure.
First-Order Tensors (Vectors): Quantities that have both magnitude and direction, e.g.:
velocity, force. The first-order tensor is symbolized with a boldface letter and by an arrow
at the top part of the vector, i.e.:
r
.
Second-Order Tensors: Quantities that have magnitude and two directions, e.g. stress and
strain. The second-order and higher-order tensors are symbolized with a boldface letter.
In the first part of this chapter we will study several tools to manage tensors (scalars,
vectors, second-order tensors, and higher-order tensors) without heeding their dependence

Tensors
1
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
10
on space and time. At the end of the chapter we will introduce tensor fields and some field
operators which can be used to interpret these fields.
In this textbook we will work indiscriminately with the following notations: tensorial,
indicial, and matricial. Additionally, when the tensors are symmetrical, it is also possible to
represent their components using the Voigt notation.
1.2 Algebraic Operations with Vectors
There now follows a brief review of vectors, in the Euclidean vector space ( ) E , so that we
may become acquainted with the nomenclature used in this textbook.
Addition: Let a
r
, b
r
be arbitrary vectors, we can show the sum of adding them, (see Figure
1.1 (a)), with a new vector ( c
r
) thus defined as:
a b b a c
r
r r
r r
+ = + =
(1.1)






Figure 1.1: Addition and subtraction of vectors.
Subtraction: The subtraction between two arbitrary vectors ( a
r
, b
r
), (see Figure 1.1 (b)), is given
as follows:
b a d
r
r
r
=
(1.2)
Considering three vectors a
r
, b
r
and c
r
the following properties are satisfied:
c b a c b a c b a
r
r
r r
r
r r
r
r
+ + = + + = + + ) ( ) ( (1.3)
Scalar multiplication: Let a
r
be a vector, we can define the scalar multiplication with a
r
.
The product of this operation is another vector with the same direction of a
r
, and whose
length and orientation is defined with the scalar as shown in Figure 1.2.






Figure 1.2: Scalar multiplication.
c
r

a
r

b
r

c
r

a
r

b
r

b
r

d
r

a) b)
a
r

1 > 0 < 1 =
a
r

a
r

a
r

1 0 < <
a
r

a
r
a
r

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
11
Scalar Product: The Scalar Product (also known as the dot product or inner product) of two
vectors a
r
, b
r
, denoted by b a
r
r
, is defined as follows:
= = cos b a b a
r
r
r
r
(1.4)
where is the angle between the two vectors, (see Figure 1.3(a)), and represents the
Euclidean norm (or magnitude) of . The result of the operation (1.4) is a scalar.
Moreover, we can conclude that a b b a
r
r r
r
= . The expression (1.4) is also true when b a
r
r
= ,
therefore:
a a a a a a a a a a a
r r r r r r r r r r r
= = =
=
2
0
cos (1.5)
Hence, the norm of a vector is a a a
r r r
= .
Unit Vector: A unit vector, associated with the a
r
-direction, is shown with a a

, which has
the same direction and orientation of a
r
. In this textbook, the hat symbol (
) denotes a
unit vector. Thus, the unit vector, a

, codirectional with a
r
, is defined as:
a
a
a r
r
=


(1.6)
where a
r
represents the norm (magnitude) of a
r
. If a

is the unit vector, then the


following must be true:
1

= a (1.7)
Zero Vector (or Null Vector): The zero vector is represented by a:
r
0 (1.8)
Projection Vector: The projection vector of a
r
onto b
r
, (see Figure 1.3(b)), is defined as:
b a a proj
b
b

r r
r
r
proj =
Projection vector of a
r
onto b
r

(1.9)
where a proj
b
r
r
is the projection of a
r
onto b
r
, and b

is the unit vector associated with the


b
r
-direction. The magnitude of a proj
b
r
r
is obtained by means of the scalar product:
b a a proj
b

=
r r
r
Projection of a
r
onto b
r

(1.10)
So, taking into account the definition of the unit vector, we obtain:
b
b a
a proj
b
r
r
r
r
r

=
(1.11)
Then, the projection vector, a proj
b
r
r
, can be calculated by:
{
b
b
b a
b
b
b
b a
b
b
b a
a proj
b
r
r
r
r
r
r
r
r
r
r
r
r
r
r

2
scalar

= = =
(1.12)

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
12








Figure 1.3: Scalar product and projection vector.
Orthogonality between vectors: Two vectors a
r
and b
r
are orthogonal if the scalar product
between them is zero, i.e.:
0 = b a
r
r

(1.13)
Vector Product (or Cross Product): The vector product of two vectors, a
r
, b
r
, results in
another vector c
r
, which is perpendicular to the plane defined by the two input vectors,
(see Figure 1.4). The vector product has the following characteristics:
Representation:
a b b a c
r
r r
r r
= =
(1.14)
The vector c
r
is orthogonal to the vectors a
r
and b
r
, thus:
0 = = c b c a
r
r
r r

(1.15)
The magnitude of c
r
is defined by the formula:
= sin b a c
r
r r
(1.16)
where measures the smallest angle between a
r
and b
r
, (see Figure 1.4).
The magnitude of the vector product b a
r
r
is geometrically expressed as the area of the
parallelogram defined by the two vectors, (see Figure 1.4):
b a
r
r
= A (1.17)
Therefore, the triangle area defined by the points OCD, (see Figure 1.4 (a)), is:
b a
r
r
=
2
1
T
A (1.18)
If a
r
and b
r
are linearly dependent, i.e. b a
r
r
= with denoting a scalar, the vector product
of two linearly dependent vectors becomes a zero vector, 0 b b b a
r r r r
r
= = .
Scalar Triple Product (or Mixed Product): Let a
r
, b
r
, c
r
be arbitrary vectors, we can
define the scalar triple product as:
( ) ( ) ( )
( ) ( ) ( ) a b c c a b b c a
b a c a c b c b a
r
r
r r r
r r
r r
r
r r r r
r
r
r
r
= = =
= = =


V
V
(1.19)
a) Scalar product b) Projection vector

a
r

b
r


a
r

b
r

.
a
b
r
r
proj
0
= cos b a b a
r
r
r
r

b


1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
13
where the scalar V represents the volume of the parallelepiped defined by c b a
r
r
r
, , , (see
Figure 1.5).








Figure 1.4: Vector product.
If two vectors are linearly dependent then, the scalar triple product is zero, i.e.:
( ) 0 a b a
r
r
r
r
= (1.20)
Let a
r
, b
r
, c
r
, d
r
, be vectors and , be scalars, the following property is satisfied:
) ( ) ( ) ( ) ( d c b d c a d c b a
r
r
r r
r r
r
r
r
r
+ = + (1.21)
NOTE: Some authors represent the scalar triple product as, ( ) c b a c b a
r
r
r r
r
r
] , , [ ,
( ) a c b a c b
r r
r
r r
r
] , , [ , ( ) b a c b a c
r
r r
r
r r
] , , [ .








Figure 1.5: Scalar triple product.
Vector Triple Product: Let a
r
, b
r
, c
r
be vectors, we can define the vector triple product as
( ) c b a w
r
r
r r
= . Then, we can demonstrate that the following relationships to be true:
( ) ( ) ( )
( ) ( )c b a b c a
a b c b a c c b a w
r
r
r
r
r r
r
r
r
r
r r r
r
r r
=
= = =
(1.22)
whereby it is clear that the result of the vector triple product is another vector w
r
, belonging to
the plane
1
formed by the vectors b
r
and c
r
, (see Figure 1.6).
a
r

b
r


a
r
b
r

.
.
V
c
r

V Scalar triple product
b a c
r
r r
=
a
r

b
r


a b c
r
r
r
=
.
.
a
r

b
r


.
.
A

T
A
O
C
D
a) b)
c
r

O
C
D
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
14













Figure 1.6: Vector triple product.
Problem 1.1: Let a
r
and b
r
be arbitrary vectors. Prove that the following relationship is
true:
( ) ( ) ( )( ) ( )
2
b a b b a a b a b a
r
r
r r
r r
r
r
r
r
=
Solution:
( ) ( )
( )
( )
( )
( )
( )( ) ( )
2
2
2
2
2 2
2
2
2
2
2
2
2
2
2
2
2
2
2
2

cos
cos
cos 1
sin
sin
b a b b a a
b a b a
b a b a
b a b a
b a
b a
b a
b a b a b a
r
r
r r
r r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r
r

=
=
=
=
=
=
=
=

Linear Transformation
Let u
r
and v
r
be arbitrary vectors, and be a scalar, we can state F is a linear
transformation if the following is true:
) ( ) ( ) ( v u v u
r r r r
F F F + = +
) ( ) ( u u
r r
F F =

1


2

a
r

b
r

c
r

c b
r
r


1
- plane defined by b
r
, c
r


2
- plane defined by a
r
, c b
r
r

w
r

w
r
belonging to the plane
1

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
15
Problem 1.2: Given the following functions = E ) ( and
2
2
1
) ( = E , demonstrate
whether these functions show a linear transformation or not.
Solution:
[ ] ) ( ) ( ) (
2 1 2 1 2 1 2 1
+ = + = + = + E E E (linear transformation)














The function
2
2
1
) ( = E does not show a linear transformation because the condition
) ( ) ( ) (
2 1 2 1
+ = + has not been satisfied:
[ ] [ ]
) ( ) ( ) ( ) (
2
2
1
2
1
2
1
2
2
1
2
1
) (
2 1 2 1 2 1
2 1
2
2
2
1
2
2 2 1
2
1
2
2 1 2 1
+ + + =
+ + = + + = + = +

E
E E E E E


















) (

2 1
+
2

1

) (
2

) (
1

) ( ) ( ) (
2 1 2 1
+ = +
2 1
+ 1

2

) (
) (
2 1
+
) (
2

) (
1

) ( ) (
2 1
+
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
16
1.3 Coordinate Systems
A tensor, which has physical meanings, must be independent of the adopted coordinate
system. Sometimes for reasons of convenience, we need to represent a tensor in a specific
coordinate system, hence, we have the concept of tensor components, (see Figure 1.7).










Figure 1.7: Tensor components.
Let a
r
be a first-order tensor (vector) as shown in Figure 1.8 (a), the tensor representation
in a general coordinate system, defined as
3 2 1
, , , is made up of its components
(
3 2 1
a a a , , ), (see Figure 1.8 (b)). Some examples of coordinate system are: the Cartesian
coordinate system; the cylindrical coordinate system; and the spherical coordinate system.






Figure 1.8: Vector representation in a general coordinate system.
1.3.1 Cartesian Coordinate System
The Cartesian coordinate system is defined by three unit vectors: i

, j

, k

, denoted by the
Cartesian basis, which make up an orthonormal basis. The orthonormal basis has the
following properties:
1. The vectors that make up this basis are unit vectors:
1

= = = k j i (1.23)
or:
1

= = = k k j j i i (1.24)
a
r

a)
a
r


1


2

3

b)
a
r
(
3 2 1
a a a , , )
TENSORS
Mathematical representation of the physical
quantities
(Independent of the coordinate system)
COMPONENTS
Tensor Representation in a
Coordinate System
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
17
2. The unit vectors ( i

, j

, k

) are mutually orthogonal, i.e.:


0

= = = i k k j j i (1.25)
3. The vector product between the vectors ( i

, j

, k

) is the following:
j i k i k j k j i

;

;

= = = (1.26)
The direction and orientation of the orthonormal basis can be obtained using the right-
hand rule as shown in Figure 1.9.
k j i

= i k j

= j i k

= (1.27)



Figure 1.9: The right-hand rule.
1.3.2 Vector Representation in the Cartesian Coordinate
System
The vector a
r
, (see Figure 1.10), in the Cartesian coordinate system, is represented by its
different components (
x
a ,
y
a ,
z
a ) and by the Cartesian bases ( i

, j

, k

) as:
k j i

z y x
a a a + + = a
r
(1.28)











Figure 1.10: Cartesian coordinate system.
Let a
r
, b
r
, c
r
be arbitrary vectors, we can describe some vector operations in the Cartesian
coordinate system, as follows:
i


x
y
z
a
r

i



x
a

y
a

z
a
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
18
The scalar product b a
r
r
becomes a scalar, which is defined in the Cartesian
system as:
) ( )

( )

(
z z y y x x z y x z y x
b a b a b a b b b a a a + + = + + + + = k j i k j i b a
r
r
(1.29)
Thus, it is true that
2
2 2 2
a a a
r r r
= + + = + + =
z y x z z y y x x
a a a a a a a a a .
NOTE: The projection of a vector onto a given direction was established in the
equation (1.10), thus defining the component concept. For example, if we want to
know the vector component along the y -direction, all we need to do is calculate:
y z y x
a a a a = + + = )

( )

(

j k j i j a
r
.
The norm of a
r
is:
2 2 2
z y x
a a a + + = a
r
(1.30)
Then, the unit vector codirectional with a
r
is:
k j i

2 2 2 2 2 2 2 2 2
z y x
z
z y x
y
z y x
x
a a a
a
a a a
a
a a a
a
+ +
+
+ +
+
+ +
= =
a
a
a r
r

(1.31)
The zero vector is:

0 0 0 = + + i j k
r
0 (1.32)
Addition: The vector sum of a
r
and b
r
is represented by:
( ) ( ) ( )k j i k j i k j i

)

( )

(
z z y y x x z y x z y x
b a b a b a b b b a a a + + + + + = + + + + + = + b a
r
r
(1.33)
Subtraction: The difference between a
r
and b
r
is:
( ) ( ) ( )k j i k j i k j i

)

( )

(
z z y y x x z y x z y x
b a b a b a b b b a a a + + = + + + + = b a
r
r
(1.34)
Scalar multiplication: The resulting vector defined by a
r
is:
k j i

z y x
a a a + + = a
r
(1.35)
The vector product ( b a
r
r
) is evaluated as:
k j i
k j i
k j i

) (

) (

) (


x y y x x z z x y z z y
y x
y x
z x
z x
z y
z y
z y x
z y x
b a b a b a b a b a b a
b b
a a
b b
a a
b b
a a
b b b
a a a
+ =
+ = = = b a c
r
r r
(1.36)
where the symbol ) ( det denotes the matrix determinant.
The scalar triple product ] c b a [
r
r
r
, , is the determinant of the 3 by 3 matrix, defined
as:

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
19
( ) ( ) ( )
( ) ( ) ( )
x y y x z x z z x y y z z y x
y x
y x
z
z x
z x
y
z y
z y
x
z y x
z y x
z y x
c b c b a c b c b a c b c b a
c c
b b
a
c c
b b
a
c c
b b
a
c c c
b b b
a a a
V
+ =
+ =
= = = = b a c a c b c b a c b a
r
r r r r
r
r
r
r r
r
r
) , , (

(1.37)
The vector triple product made up of the vectors ( c b a
r
r
r
, , ) is obtained, in the
Cartesian coordinate system, as:
( ) ( ) ( )
( ) ( ) ( )k j i


2 1 2 1 2 1 z z y y x x
c b c b c b + + =
= c b a b c a c b a
r
r
r
r
r r r
r
r
(1.38)
where
z z y y x x
c a c a c a + + = = c a
r r
1
, and
z z y y x x
b a b a b a + + = = b a
r
r
2
.
Problem 1.3: Consider the points: ( ) 1 , 3 , 1 A , ( ) 1 , 1 , 2 B , ( ) 3 , 1 , 0 C and ( ) 4 , 2 , 1 D , defined in the
Cartesian coordinate system.
1) Find the parallelogram area defined by

AB and

AC ; 2) Find the volume of the


parallelepiped defined by

AB,

AC and

AD ; 3) Find the projection vector of

AB onto

BC .
Solution:
1) Firstly we calculate the vectors

AB and

AC :
( ) ( ) k j i k j i k j i

0

2 + = + + + = = =

OA OB AB a
r

( ) ( ) k j i k j i k j i

2

0 + = + + + + = = =

OA OC AC b
r

With reference to the equation (1.36) we can evaluate the vector product as follows:
k j i
k j i

) 6 (

) 8 (
2 2 1
0 4 1

+ =

= b a
r
r

Then, the parallelogram area can be obtained using definition (1.19), thus:
104 ) 6 ( ) 2 ( ) 8 (
2 2 2
= + + = = b a
r
r
A
2) Next, we can evaluate the vector

AD as:
( ) ( ) k j i k j i k j i

3

1 + = + + + + = = =

OA OD AD c
r

and using the equation (1.37) we can obtain the volume of the parallelepiped:
( ) ( ) ( )
16 18 2 0

0 ) , , (
= + =
+ = = k j i k j i b a c c b a
r
r r r
r
r
V

3) The

BC vector can be calculated as:


( ) ( ) k j i k j i k j i

2

0 + + = + + + = =

OB OC BC
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
20
Hence, it is possible to evaluate the projection vector of

AB onto

BC , (see equation
(1.12)), as:
( ) ( )
( ) ( )
( )
( )
( )
( ) k j i k j i
k j i
k j i k j i
k j i k j i

3
5

3
5

3
5

2
4 4 4
0 8 2

2
2
= + +
+ +
+
=
+ +
+ + + +
+ + +
= =

BC
BC BC
AB BC
AB
BC
BC
43 42 1
proj

1.3.3 Einstein Summation Convention (Einstein Notation)
As we saw in equation (1.28) a
r
in the Cartesian coordinate system was defined as:
k j i

z y x
a a a + + = a
r
(1.39)
Said expression can be rewritten as:
3 3 2 2 1 1

e e e a a a a + + =
r
(1.40)
where we have considered that:
x
a a
1
,
y
a a
2
,
z
a a
3
, i

1
e , j

2
e , k

3
e , (see
Figure 1.11). In this way we can express equation (1.40) by means of the summation
symbol as:

=
= + + =
3
1
3 3 2 2 1 1

i
i i
e e e e a a a a a
r
(1.41)
Then, we introduce the summation convention, according to which the repeated indices
indicate summation. So, equation (1.41) can be represented as follows:
) 3 , 2 , 1 (

3 3 2 2 1 1
= = + + = i
i i
e e e e a a a a a
r

) 3 , 2 , 1 (

= = i
i i
e a a
r

(1.42)
NOTE: The summation notation was introduced by Albert Einstein in 1916, which led to
the indicial notation.
1.4 Indicial Notation
Using indicial notation, the three axes of the coordinate system are designated by the letter
x with a subscript. So,
i
x is not a single value but i values, i.e.
1
x ,
2
x ,
3
x (if 3 , 2 , 1 = i )
where these values
1
x ,
2
x ,
3
x correspond to the axes x , y , z , respectively.
Let a
r
be a vector represented in the Cartesian coordinate system as:
3 3 2 2 1 1

e e e a a a a + + =
r
(1.43)
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
21
where the orthonormal basis is represented by
1

e ,
2

e ,
3

e , (see Figure 1.11), and


1
a ,
2
a ,
3
a are the vector components. In indicial notation the vector components are represented
by
i
a . If the range of the subscript is not indicated, we assume that 1,2,3 show these
values. Therefore, the vector components are represented as:
(
(
(

= =
3
2
1
) (
a
a
a
a
i i
a
r
(1.44)












Figure 1.11: Vector representation in the Cartesian coordinate system.
Unit vector components: Let a
r
be a vector, the normalized vector a

is defined as:
1

with

= = a
a
a
a r
r

(1.45)
whose components are:
) 3 , 2 , 1 , , (

2
3
2
2
2
1
= = =
+ +
= k j i
k k
i
j j
i i
i
a a
a
a a
a
a a a
a
a
(1.46)
In light of the previous equation we can emphasize two types of indices:
The free index (live index) is that which only appears once in a term of the expression. In the
above equation the free index is the ( i ). The number of the free index indicates
the tensor order.
The dummy index (summation index) is that which is repeated only twice in a term of the
expression, and indicates summation. In the above equation (1.46) the
dummy index is the ( j ), or the ( k ) index.
OBS.: An index in a term of an expression can only appear once or twice. If it
appears more times, then a large error has occurred.

2
x y

1
x x

3
x z
a
r


1

e i
2

e j

3

e k

1
a a
x


2
a a
y


3
a a
z

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
22
Scalar product: Using definitions (1.4) and (1.29), we can express the scalar product
( b a
r
r
) as follows:
) 3 , 2 , 1 , ( cos
3 3 2 2 1 1
= = = + + = = = j i
j j i i
b a b a b a b a b a b a b a
r
r
r
r
(1.47)


Problem 1.4: Rewrite the following equations using indicial notation:

1)
3 3 3 3 2 2 3 1 1
x x a x x a x x a + +
Solution: ) 3 , 2 , 1 (
3
= i x x a
i i

2)
2 2 1 1
x x x x +
Solution: ) 2 , 1 ( = i x x
i i

3)

= + +
= + +
= + +
z
y
x
b z a y a x a
b z a y a x a
b z a y a x a
33 32 31
23 22 21
13 12 11

Solution:

= + +
= + +
= + +
3 3 33 2 32 1 31
2 3 23 2 22 1 21
1 3 13 2 12 1 11
b x a x a x a
b x a x a x a
b x a x a x a

j index
dummy

=
=
=
3 3
2 2
1 1
b x a
b x a
b x a
j j
j j
j j

i index
free


i j ij
b x a =
As we can appreciate in this problem, the use of the indicial notation means that the
equation becomes very concise. In many cases, if algebraic operation do not use indicial or
tensorial notation they become almost impossible to deal with due to the large number of
terms involved.
Problem 1.5: Expand the equation: ) 3 , 2 , 1 , ( = j i x x A
j i ij

Solution: The indices j i, are dummy indices, and indicate index summation and there is no
free index in the expression
j i ij
x x A , therefore the result is a scalar. So, we expand first the
dummy index i and later the index j to obtain:
43 42 1 43 42 1 43 42 1
3 3 33
2 3 32
1 3 31
3 3
3 2 23
2 2 22
1 2 21
2 2
3 1 13
2 1 12
1 1 11
1 1
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A
x x A x x A
j j j j j j
i expanding
j i ij
+
+
+
+
+
+
+
+


Rearranging the terms we obtain:
3 3 33 2 3 32 1 3 31 3 2 23
2 2 22 1 2 21 3 1 13 2 1 12 1 1 11
x x A x x A x x A x x A
x x A x x A x x A x x A x x A x x A
j i ij
+ + +
+ + + + + =

1.4.1 Some Operators
1.4.1.1 Kronecker Delta
The Kronecker delta
ij
is defined as follows:
e
x
p
a
n
d
i
n
g


j

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
23

=
=
j i iff
j i iff
ij
0
1
(1.48)
Also note that the scalar product of the orthonormal basis
j i
e e

is equal to 1 if j i = and
equal to 0 if j i . Hence,
j i
e e

can be expressed in matrix form as:
ij j i
=
(
(
(

=
(
(
(

1 0 0
0 1 0
0 0 1




3 3 2 3 1 3
3 2 2 2 1 2
3 1 2 1 1 1
e e e e e e
e e e e e e
e e e e e e
e e (1.49)
An interesting property of the Kronecker delta is shown in the following example. Let
i
V
be the components of the vector V
r
, therefore:
3 3 2 2 1 1
V V V V
j j j i ij
+ + =
(1.50)
As ) 3 , 2 , 1 ( = j is a free index, we have three values to be calculated, namely:
j i ij
i ij
i ij
i ij
V V
V V V V V j
V V V V V j
V V V V V j
=

= + + = =
= + + = =
= + + = =




3 3 33 2 23 1 13
2 3 32 2 22 1 12
1 3 31 2 21 1 11
3
2
1
(1.51)
That is, in the presence of the Kronecker delta symbol we replace the repeated index as
follows:
j i j i
V V

=
(1.52)
For this reason, the Kronecker delta is often called the substitution operator.
Other examples using the Kronecker delta are presented below:
jk ik ij
A A = , 3
33 22 11
= + + = = =
jj ii ji ij
,
33 22 11
a a a a a
ii ji ji
+ + = =
(1.53)
To obtain the components of the vector a
r
in the coordinate system represented by
i
e

, it
is sufficient to obtain the scalar product with a
r
and
i
e

, i.e.
i pi p i p p i
a a a = = = e e e a

r
.
With that, it is also possible to represent the vector as:
i i i i
e e a e a

)

= =
r r
a (1.54)
Problem 1.6: Solve the following equations:
1)
jj ii

Solution: ( )( ) 9 3 3
33 22 11 33 22 11
= = + + + + =
jj ii

2)
1 1

Solution: 1
11 1 1 1 1
= = =


NOTE: Note that the following algebraic operation is incorrect 1 3
11 1 1
= =

,
since what must be replaced is the repeated index, not the number
1.4.1.2 Permutation Symbol
The permutation symbol
ijk
(also known as Levi-Civita symbol or alternating symbol) is defined as:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
24
{ }
{ }

= = =

+
=
) ( ) ( ) (
) 3 , 1 , 2 ( ), 1 , 2 , 3 ( ), 2 , 3 , 1 ( ) , , ( 1
) 2 , 1 , 3 ( ), 1 , 3 , 2 ( ), 3 , 2 , 1 ( ) , , ( 1
0 k i or k j or j i if
k j i if
k j i if
: i.e. cases remaining the for
ijk
(1.55)
NOTE:
ijk
are the components of the Levi-Civita pseudo-tensor, which will be introduced
later on.
The values of
ijk
can be easily memorized using the mnemonic device shown in Figure
1.12(a), in which if the index values are arranged in a clockwise direction, the value of
ijk

is equal to 1, if not it has the value of 1 . In the same way we can use this mnemonic
device to switch indices, (see Figure 1.12(b)).








Figure 1.12: Mnemonic device for the permutation symbol.
Another way to express the permutation symbol is by means of its indices:
) )( )( (
2
1
i k k j j i
ijk
= (1.56)
Using both the definition seen in (1.55) and Figure 1.12 (b), it is possible to verify that the
following relations are valid:
kji jik ikj ijk
kij jki ijk


= = =
= =
(1.57)
Using the Kronecker delta property, we can state that:
( ) ( ) ( )
j i j i k k i k i j k j k j i
k j i k j i k j i k j i k j i k j i
nk mj li lmn ijk
2 3 3 2 1 2 3 3 2 1 2 3 3 2 1
1 2 3 1 3 2 2 1 3 3 1 2 2 3 1 3 2 1



+ =
+ + =
=

(1.58)
The above equation can be represented by means of the following determinant:
k k k
j j j
i i i
k j i
k j i
k j i
ijk
3 2 1
3 2 1
3 2 1
3 3 3
2 2 2
1 1 1






= = (1.59)
After which, the term
pqr ijk
can be evaluated as follows:
r q p
r q p
r q p
k k k
j j j
i i i
pqr ijk
3 3 3
2 2 2
1 1 1
3 2 1
3 2 1
3 2 1






= (1.60)
1
2
3
1 =
ijk


1 =
ijk


a)
i
j
k
b)

kij jki ijk
= =

ikj ijk
=


jik
kji
ikj ijk


=
=
=

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
25
Taking into account that ) ( ) ( ) ( B A AB det det det = , where ) ( det is the determinant
of the matrix , the equation (1.60) can be rewritten as:

(
(
(

(
(
(

=
3 3 3
2 2 2
1 1 1
3 2 1
3 2 1
3 2 1
r q p
r q p
r q p
k k k
j j j
i i i
pqr ijk







kr kq kp
jr jq jp
ir iq ip
pqr ijk



= (1.61)
The term
ip
was obtained by means of the operation
mp mi p i p i p i
= + +
3 3 2 2 1 1

and
ip mp mi
= , the other terms were obtained in a similar fashion.
For the special exception when k r = , the equation (1.61) is reduced to:

3
=
kq kp
jk jq jp
ik iq ip
pqr ijk



3 , 2 , 1 , , , , = = q p k j i
jp iq jq ip pqk ijk
(1.62)
Problem 1.7: a) Prove the following is true
ip pjk ijk
2 = and 6 =
ijk ijk
. b) Obtain the
numerical value of
i k j ijk 1 3 2
.
Solution: a) Using the equation in (1.62), i.e.
jp iq jq ip pqk ijk
= , and by substituting q
for j , we obtain:
ip ip ip jp ij jj ip pjk ijk
2 3 = = =
Based on the above result, it is straight forward to check that:
6 2 = =
ii ijk ijk

b) 1
123 1 3 2
= =
i k j ijk

The vector product of two vectors ( b a
r
r
) leads to a new vector c
r
, defined in
(1.36), and the components of c
r
, in Cartesian system, are given by:
3
3
1 2 2 1 2
2
3 1 1 3 1
1
2 3 3 2
3 2 1
3 2 1
3 2 1

) (

) (

) (

e e e
e e e
b a c
43 42 1 43 42 1 4 43 4 42 1
r
r r
c c c
b a b a b a b a b a b a
b b b
a a a
+ + =
= =
(1.63)
Using the definition of the permutation symbol
ijk
, defined in (1.55), we can express the
components of c
r
as follows:
k j ijk i
k j jk
k j jk
k j jk
b a c
b a b a b a c
b a b a b a c
b a b a b a c

= + =
= + =
= + =
3 1 2 321 2 1 312 3
2 3 1 213 1 3 231 2
1 2 3 132 3 2 123 1
(1.64)
Then, the vector product ( b a
r
r
) can be represented by means of the permutation symbol
as:
i jki k j i ijk k j k j k j
i ijk k j k k j j
i k j ijk
e e e e
e e e
e b a

)

(


b a b a b a
b a b a
b a
= =
=
=
r
r
(1.65)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
26
Therefore, we can also conclude that the following relationship is valid:
i ijk k j
e e e

)

( = (1.66)
The permutation symbol and the orthonormal basis can be interrelated using the triple
scalar product as follows:
( ) ( )
ijk k j i mk ijm k m ijm k j i m ijm j i
= = = = e e e e e e e e e e e

(1.67)
The triple scalar product made up of the vectors c b a
r
r
r
, , is expressed by:
( ) ( ) ( )
k j i ijk k j i k j i k k j j i i
c b a c b a c b a = = = = e e e e e e c b a

r
r
r
(1.68)
( ) ) 3 , 2 , 1 , , ( = = = k j i
k j i ijk
c b a c b a
r
r
r

(1.69)
or
( ) ( ) ( )
3 2 1
3 2 1
3 2 1
c c c
b b b
a a a
= = = = b a c a c b c b a
r
r r r r
r
r
r
r
(1.70)
Starting from the equation (1.69) we can prove the following are true:
( ) ( ) ( ) b a c a c b c b a
r
r r r r
r
r
r
r
= = :
( )
( )
( )
( )
( )
( ) ] , , [
] , , [
] , , [
] , , [
] , , [
] , , [
a b c a b c
c a b c a b
b c a b c a
b a c b a c
a c b a c b
c b a c b a
r
r
r r
r
r
r r
r
r r
r
r
r r
r
r r
r
r r
r
r r
r r
r
r r
r
r
r
r r
r
r
= =
= =
= =
= =
= =
=

k j i kji
k j i jik
k j i ikj
k j i kij
k j i jki
k j i ijk
c b a
c b a
c b a
c b a
c b a
c b a

(1.71)
where we take into account the property of the permutation symbol as given in (1.57).
Problem 1.8: Rewrite the expression ( ) ( ) d c b a
r
r
r
r
without using the vector product
symbol.
Solution: The vector product ( ) b a
r
r
can be expressed as
( )
i k j ijk k k j j
e e e b a

b a b a = =
r
r
. Likewise, it is possible to express ( ) d c
r
r
as
( )
n m l nlm
e d c

d c =
r
r
, thus:
( ) ( )
m l k j ilm ijk in m l k j nlm ijk
n i m l k j nlm ijk n m l nlm i k j ijk
d c b a d c b a
d c b a d c b a


= =
= =

e e e e d c b a

)

( )

r
r
r
r

Taking into account that
lmi jki ilm ijk
= (see equation (1.57)) and by applying the
equation (1.62), i.e.:
ilm jki kl jm km jl lmi jki
= = , we obtain:
( )
m l l m m l m l m l k j kl jm km jl m l k j ilm ijk
d c b a d c b a d c b a d c b a = =
Since ( ) c a
r r
=
l l
c a and ( ) d b
r r
=
m m
d b holds true, we can conclude that:
( ) ( ) ( )( ) ( )( ) c b d a d b c a d c b a
r
r r
r
r r
r r
r
r
r
r
=
Therefore, it is also valid when c a
r r
= and d b
r r
= , thus:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
27
( ) ( ) ( )( ) ( )( ) ( )
2 2
2
2
b a b a a b b a b b a a b a b a b a
r
r
r
r r
r r
r
r r
r r
r
r
r
r
r
r
= = =
which is the same equation obtained in Problem 1.1.
Problem 1.9: Prove that ( ) ( ) [ ] [ ] ) ( ) ( b a c d b a d c d c b a
r
r r
r r
r
r
r
r
r
r
r
=
Solution: Expressing the correct equality term in indicial notation we obtain:
[ ] [ ] ( ) [ ] ( ) [ ]
k j ijk i p k j ijk i p
p
b a c d b a d c ) ( ) ( =
)
`

b a c d b a d c
r
r r
r r
r
r
r

( )
p i i p
k
j ijk p i k j ijk i p k j ijk
d c d c b a d c b a d c b a
Using the Kronecker delta the above equation becomes:
( ) ( ) ( )
np im ni pm n m k j ijk np n m im ni n m pm
k
j ijk
d c b a d c d c b a
and by applying the equation
mnl pil np im ni pm
= , (see eq. (1.62)), the above equation
can be rewritten as follows:
( ) ( ) ( )( ) [ ]
n m mnl k j ijk pil mnl pil n m k j ijk
d c b a d c b a
Since
k j ijk
b a and
n m mnl
d c represent the components of ( ) b a
r
r
and ( ) d c
r
r
,
respectively, we can conclude that:
( )( ) [ ] ( ) ( ) [ ]
p n m mnl k j ijk pil
d c b a
r
r
r
r
= d c b a
Problem 1.10: Let a
r
, b
r
, c
r
be linearly independent vectors, and v
r
be a vector,
demonstrate that:
0 c b a v
r
r
r
r r
+ + =
where the scalars , , are given by:
r q p pqr
k j i ijk
r q p pqr
k j i ijk
r q p pqr
k j i ijk
c b a
v b a
c b a
c v a
c b a
c b v

= = = ; ;
Solution: The scalar product made up of v
r
and ( c b
r
r
) becomes:
43 42 1
r
r
r
43 42 1
r
r r
r
r
r r
r
r
0 0
) ( ) ( ) ( ) (
= =
+ + = c b c c b b c b a c b v
) (
) (
c b a
c b v
r
r
r
r
r
r


which is the same as:
r q p pqr
k j i ijk
c b a
c b v
c b a
c b a
c b a
c b v
c b v
c b v
c c c
b b b
a a a
c c c
b b b
v v v

= = =
3 3 3
2 2 2
1 1 1
3 3 3
2 2 2
1 1 1
3 2 1
3 2 1
3 2 1
3 2 1
3 2 1
3 2 1


One can obtain the parameters and in a similar fashion.
Problem 1.11: Prove the relationship given in (1.38) is valid, i.e.:
( ) ( ) ( )c b a b c a c b a
r
r
r
r
r r r
r
r
= .
Solution: Taking into account that ( )
k j ijk i i
c b = = c b d
r
r r
) ( and that ( )
k j qjk q
c b = d a
r
r
, we
obtain:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
28
( ) [ ]
( )
( ) ( ) b a c a
c b a
r
r r r
r
r
r
= =
= =
= = =
q q q j j k q k
k j s sj qk k j s sk qj k j s sj qk sk qj
k j s jki qsi k j s ijk qsi k j ijk s qsi q
c b c b a c b a
c b a c b a c b a
c b a c b a c b a

) (

( ) [ ] ( ) ( ) [ ]
q q
b a c c a b c b a
r
r r r r
r
r
r
r
=
1.5 Algebraic Operations with Tensors
1.5.1 Dyadic
The tensor product, made up of two vectors v
r
and u
r
, becomes a dyad, which is a particular
case of a second-order tensor. The dyad is represented by:
A v u v u =
r r r r
(1.72)
where the operator denotes the tensor product. Then, we define a dyadic as a linear
combination of dyads. Furthermore, as we will see later, any tensor can be represented by
means of a linear combination of dyads, (see Holzapfel (2000)).
The tensor product has the following properties:
1. ) ( ) ( ) ( x v u x v u x v u
r r r r r r r r r
=
(1.73)
2. w u v u w v u
r r r r r r r
+ = + ) ( (1.74)
3.
[ ] [ ] ) ( ) (
) ( ) ( ) (
x r w x u v
x r w x u v x r w u v
r r r r r r
r r r r r r r r r r r


+ =
+ = +


(1.75)
where and are scalars. By definition, the dyad does not contain the commutative
property, i.e., u v v u
r r r r
.
The equation (1.72) can also be expressed in the Cartesian system as:
)

(
)

(
)

( )

(
j i ij
j i j i
j j i i
e e
e e
e e v u A
=
=
= =
A
v u
v u
r r
) 3 , 2 , 1 , ( = j i (1.76)
{
{ 43 42 1
basis
j i
components
ij
Tensor
e e A

= A
) 3 , 2 , 1 , ( = j i (1.77)
In this textbook, the components of a second-order tensor can be represented in different
ways, namely:
ij j i ij ij
components
A v u = = =
=

) ( ) ( v u A
v u A
r r
43 42 1
r r
(1.78)
These components are explicitly expressed in matrix form as:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
29
(
(
(

= = =
33 32 31
23 22 21
13 12 11
) (
A A A
A A A
A A A
A A
ij ij
A (1.79)
It is easy to identify the tensor order by the number of free indices in the tensor
components, i.e.:
Second-order tensor
Third-order tensor
Fourth-order tensor l k j i ijkl
k j i ijk
j i ij
e e e e
e e e T
e e U



=
=
=
I I
T
U

) 3 , 2 , 1 , , , ( = l k j i (1.80)

Problem 1.12: Define the order of the tensors represented by their Cartesian components:
i
v ,
ijk
,
ijj
F ,
ij
,
ijkl
C ,
ij
. Determine the number of components in tensor C .
Solution: The order of the tensor is given by the number of free indices, so it follows that:
First-order tensor (vector): v
r
, F
r
; Second-order tensor: , ; Third-order tensor: ;
Fourth-order tensor: C
The number of tensor components is given by the maximum index range value, i.e.
3 , 2 , 1 , , , = l k j i , to the power of the number of free indices which is equal to 4 in the case of
ijkl
C . Thus, the number of independent components in C is given by:
( ) ( ) ( ) ( ) 81 3 3 3 3 3
4
= = = = = = l k j i
The fourth-order tensor
ijkl
C has 81 components.
Let A and B be second-order tensors, we can then define some algebraic operations
including:
Addition: The sum of two tensors of same order is a new tensor defined as follows:
A B B A C + = + = (1.81)
The components of C are represented by:
ij ij
) ( ) ( B A C + = or
ij ij ij
B A C + =
(1.82)
or, in matrix notation as:
B A C + = (1.83)
Multiplication of a tensor by a scalar: The multiplication of a second-order tensor
( A ) by a scalar ( ) is defined by a new tensor D , so that:
ij ij
components in
) ( ) (

A D A D = = (1.84)
or, in matrix form:
OBS.: The tensor order is given by the number of free indices in its components.
OBS.: The number of tensor components is given by
n
a , where the base a is the
maximum value in the index range, and the exponent n is the number of the free
index.
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
30
(
(
(




=
(
(
(

=
33 32 31
23 22 21
13 12 11
33 32 31
23 22 21
13 12 11
A A A
A A A
A A A
A A A
A A A
A A A
A A (1.85)
It is also true that:
) ( ) ( v A v A
r r
= (1.86)
for any vector v
r
.
Scalar Product (or Dot Product): The scalar product (also known as single
contraction) between a second-order tensor A and a vector x
r
is another vector
(first-order tensor) y
r
, defined as:

j j
j
j
k jk
j kl l jk
l l k j jk
e
e
e
e e e
x A y

( )

(
y
x A
x A
x A
y
=
=
=
=
=

3 2 1
r r

(1.87)
The scalar product between two second-order tensors A and B is another second-order
tensor, that verifies: A B B A :


l i il
l i kl ik
l i jk kl ij
l k kl j i ij
e e
e e
e e
e e e e B A C



)

( )

(
=
=
=
= =
C
B A
B A
B A
3 2 1
AB


l i il
l i kl ik
l i jk kl ij
l k kl j i ij
e e
e e
e e
e e e e A B D



)

( )

(
=
=
=
= =
D
A B
A B
A B
3 2 1
BA

(1.88)
It also satisfies the following properties:
C A B A C B A + = + ) ( ; C B A C B A = ) ( ) (
(1.89)
The powers of second-order tensors
The scalar product allows us to define the power of second-order tensors, as seen below:
A A A A A A A A A 1 A = = = =
3 2 1 0
; ; ; , and so on, (1.90)
where 1 is the second-order unit tensor (also called the identity tensor).

Double Scalar Product (or Double contraction)
Consider two dyads, d c A
r
r
= and v u B
r r
= . The double contraction between them is
defined in different ways, namely: B A
:
and B A .
Double contraction : ) (
( ) ( ) ( )( ) u d v c v u d c
r
r
r r r r
r
r
= (1.91)
In components

kl


jk

jk

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
31



) (
)

( )

(
scalar
ji ij
il jk kl ij
l k kl j i ij


=
=
=
=
B A
B A
B A e e e e B A

(1.92)
The double contraction ) ( is commutative, i.e. A B B A = .
Double contraction (: ):
( ) ( ) ( )( ) v d u c v u d c B A
r
r
r r r r
r
r
= = :
:
(1.93)
The double contraction (: ) is commutative, so:
( ) ( ) ( )( ) ( )( ) B A v d u c d v c u d c v u A B
:
:
:
= = = =
r
r
r r
r
r r r
r
r r r
(1.94)
The breakdown into its components appears like this:


) (
)

( )

(
scalar
ij ij
jl ik kl ij
l k kl j i ij
=
=
=
=
B A
B A
B A

e e e e B A
: :
(1.95)
In general, B A B A
:
, however, they are equal if at least one of them is symmetric, i.e.
B A B A =
sym sym
:
or
sym sym
B A B A =
:
, so
sym sym sym sym
B A B A =
:
.
The double contraction with a third-order tensor ( S ) and a second-order tensor ( B )
becomes:
( ) ( ) ( )( )
( ) ( ) ( )( )a d v c u a d c v u S B
c u d v a v u a d c B S
r
r
r r r r
r
r r r
r r
r
r r r r r
r
r


= =
= =
:
:
:
:
(1.96)
As we can verify the result is a vector. In symbolic notation, the double contraction ( S B
:
)
is represented by:

i jk ijk i kq jp pq ijk q p pq k j i ijk
e e e e e e e

B S B S B S = =
:
(1.97)

The double contraction of a fourth-order tensor ( C ) with a second-order tensor ( ) is
defined as:

j i ij
j i kl ijkl j i lq kp pq ijkl q p pq l k j i ijkl
e e
e e e e e e e e e e


=
= = C C C
:
(1.98)
where
ij
are the components of
:
C = .

ik


jl


il


jk

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
32
Next, we express some properties of the double contraction (: ):
( )
( ) ( ) ( ) B A B A B A
C A B A C B A
A B B A
= =
+ = +
=
: : :
: : :
: :
)
)
)
c
b
a

(1.99)
where C B A , , are second-order tensors, and is a scalar.
Via the definition of the double scalar product, it is possible to obtain the components of
the second-order tensor A in the Cartesian system, i.e.:
ij lj ki kl j l k kl i j i l k kl ij
A A A A = = = = e e e e e e e e A

)

(

)

( )

( ) (
:
(1.100)
If we consider any two vectors a
r
, b
r
, and an arbitrary second-order tensor, A , we can
demonstrate that:
) (
) (

b a A
e e e e b A a
r
r
r
r
=
= = = =
:
j i ij j ij i jr pi r ij p r r j i ij p p
b a A b A a b A a b A a
(1.101)
Vector product
The vector product between a second-order tensor A and a vector x
r
is a second-order
tensor given by:
l i k ij ljk k k j i ij
e e e e e x A

)

( )

( = = x A x A
r
(1.102)
where we have used the definition (1.67), i.e.
l ljk k j
e e e

= . In Problem 1.11, we have
shown that the relation ( ) ( ) ( )c b a b c a c b a
r
r
r
r
r r r
r
r
= holds, which is also represented by
means of dyads as:
( ) [ ] ( ) [ ]
j k k j k j j k k j k k j
a b c c b c b a
r
r
r r
r
r
r
r
= = = a b c c b c b a b c a ) ( ) ( ) ( (1.103)
In the particular case when c a
r r
= we obtain:
( ) [ ]
[ ] [ ]
[ ] { }
j
p j p jp k k p j kp k jp k k
j kp p k jp p k k j k k j k k j
b a a 1 a a
a b a
r
r r r r
r
r
r
=
= =
= =
) (
) ( ) ( ) (
) ( ) ( ) ( ) (
b a a a a b a a a a
a b a b a a a b a b a a


(1.104)
Thus, the following relationships are valid:
( ) ( ) ( )
[ ] b a a 1 a a a b a
a b c c b c b a b c a c b a
r
r r r r r
r
r
r
r
r r
r
r
r
r
r
r r r
r
r


=
= =
) ( ) (
) (
(1.105)
1.5.1.1 Component Representation of a Second-Order Tensor in the
Cartesian Basis
As seen before, a vector which has 3 independent components is represented in a
Cartesian space as shown in Figure 1.11. An arbitrary second-order tensor has 9
independent components, so we would need a hyperspace to represent all its components.
Afterwards, a device is introduced to represent the second-order tensor components in the
Cartesian basis.
An arbitrary second-order tensor T is represented in the Cartesian basis by:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
33
3 3 33 2 3 32 1 3 31
3 2 23 2 2 22 1 2 21
3 1 13 2 1 12 1 1 11
3 3 2 2 1 1




e e e e e e
e e e e e e
e e e e e e
e e e e e e e e T
+ + +
+ + + +
+ + + =
+ + = =
T T T
T T T
T T T
T T T T
i i i i i i j i ij
(1.106)
Next, we calculate the projection of T onto
k
e

:
3 3 2 2 1 1

e e e e e e e e e T
k k k i ik jk i ij k j i ij k
T T T T T T + + = = = = (1.107)
thereby defining three vectors, namely:

= + + = =
= + + = =
= + + = =
=
)

(
3 33 2 23 1 13 3
)

(
3 32 2 22 1 12 2
)

(
3 31 2 21 1 11 1
3
2
1

3

2

1

e
e
e
t e e e e
t e e e e
t e e e e
e e T
r
r
r
T T T T
T T T T
T T T T
T
i i
i i
i i
i ik k
k
k
k
(1.108)
Graphical representation of these three vectors
)

(
1
e
t
r
,
)

(
2
e
t
r
,
)

(
3
e
t
r
, in the Cartesian basis, is
shown in Figure 1.13. Note also that
)

(
1
e
t
r
is the projection of T onto
1

e , [ ] 0 , 0 , 1

) 1 (
=
i
n ,
which can be verified by:
)

(
31
21
11
33 32 31
23 22 21
13 12 11
1
0
0
1
)

(
e
n T
i i
t
T
T
T
T T T
T T T
T T T
=
(
(
(

=
(
(
(

(
(
(

= (1.109)









Figure 1.13: The projection of T in the Cartesian basis.
The same result obtained in (1.109) could have been evaluated by the scalar product of T ,
given in (1.106), with the basis
1

e , i.e.:
[
]
)

(
3 31 2 21 1 11
1 3 3 33 2 3 32 1 3 31
3 2 23 2 2 22 1 2 21
3 1 13 2 1 12 1 1 11 1
1




e
t e e e
e e e e e e e
e e e e e e
e e e e e e e T
r
= + + =
+ + +
+ + + +
+ + + =

T T T
T T T
T T T
T T T
(1.110)
where we have used the orthogonality property of the basis, i.e. 1

1 1
= e e , 0

1 2
= e e ,
0

1 3
= e e . Taking into account the components are represented in matrix form, (see
Figure 1.14), we can establish that, the diagonal terms (
11
T ,
22
T ,
33
T ) are normal to the
plane defined by the unit vectors (
1

e ,
2

e ,
3

e ), hence they will be referred to as normal



1
x

2
x

3
x

)

(
2
e
t
r


)

(
3
e
t
r


)

(
1
e
t
r


2

e

3

e

1

e
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
34
components. The components displayed tangentially to the plane are called tangential
components, and correspond to the off-diagonal terms of
ij
T .














Figure 1.14: Representation of the second-order tensor components in the Cartesian
coordinate system.
NOTE: Throughout the textbook, we will use the following notations:


)

(
)

(
)

( )

(
l i jl ij
l i jk kl ij
l k kl j i ij
e e
e e
e e e e B A
=
=
=
B A
B A
B A

(1.111)

Note that the index is not repeated more than twice either in symbolic notation or in
indicial notation. Also note that the indicial notation is equivalent to the tensor notation
only when dealing with scalars, e.g. = =
ij ij
B A B A: , or
i i
b a = b a
r
r
.
1.5.2 Properties of Tensors
1.5.2.1 Tensor Transpose
Let A be a second-order tensor, the transpose of A is defined as:
)

( )

(
i j ij j i ji
T
e e e e A = = A A (1.112)
If
ij
A are the components of A , it follows that the components of the transpose are:
Tensorial notation
Symbolic notation
Cartesian basis
Indicial notation

1
x

2
x

3
x

1 11

e T

2 21

e T
3 31

e T

1 12

e T

3 32

e T

2 22

e T

3 33

e T

1 13

e T
2 23

e T

)

(
1
e
t
r


)

(
2
e
t
r


)

(
3
e
t
r


(
(
(

=
33 32 31
23 22 21
13 12 11
T T T
T T T
T T T
T
ij

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
35
ji ij
T
A = ) (A (1.113)
If v u A
r r
= , the transpose of the dyad A is given by u v A
r r
=
T
:
( )
( )
( )
( )
j i ji i j ij
T
j i ij
j i i j i j j i
T
j i j i
j j i i i i j j
T
j j i i
T
T
e e e e e e
e e e e e e
e e e e e e
u v v u A



= = =
= = =
= = =
= =
A A A
v u v u v u
u v u v v u
r r r r
(1.114)
Let A and B be second-order tensors and , be scalars, and the following
relationships are valid:
A A =
T T
) ( ;
T T T
A B A B + = + ) ( ;
T T T
B A A B = ) ( (1.115)
( ) ( )
( ) ( ) B A e e e e B A
B A e e e e B A


= = = =
= = = =
ji ij il jk kl ij l k kl i j ij
T
ji ij jk il kl ij k l kl j i ij
T
B A B A B A
B A B A B A




: :
: :

(1.116)
The transpose of the matrix A is formed by changing rows for columns and vice versa,
i.e.:
(
(
(

=
(
(
(

=
(
(
(

=
33 23 13
32 22 12
31 21 11
33 32 31
23 22 21
13 12 11
33 32 31
23 22 21
13 12 11
A A A
A A A
A A A
A A A
A A A
A A A
A A A
A A A
A A A
T
T transpose
A A (1.117)
Problem 1.13: Let A , B and C be arbitrary second-order tensors. Demonstrate that:
( ) ( ) ( ) B C A C A B C B A : : :
T T
= =
Solution: Expressing the term ( ) C B A : in indicial notation we obtain:
( ) ( )
( )
kj ik ij jq il kp pq lk ij
q l kp j i pq lk ij
q p pq k l lk j i ij
C B A C B A
C B A
C B A
= =
=
=

e e e e
e e e e e e C B A


:
: :

Note that, when we are dealing with indicial notation the position of the terms does not
matter, i.e.:
ik kj ij kj ij ik kj ik ij
B C A C A B C B A = =
We can now observe that the algebraic operation
ij ik
A B is equivalent to the components of
the second-order tensor
kj
T
) ( A B , thus,
( ) C A B A B : = =
T
kj kj
T
kj ij ik
C C A B ) ( .
Likewise, we can state that ( ) B C A :
T
ik kj ij
= B C A .
Problem 1.14: Let u
r
, v
r
be vectors and A be a second-order tensor. Show that the
following relationship holds:
u A v v A u
r r r r
=
T

Solution:
l jl j j jl l
il i jl kj k jk k il jl i
i i l j jl k k k k j l jl i i
T
u A v v A u
u A v v A u
u A v v A u
=
=
=
=



e e e e e e e e
u A v v A u

r r r r

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
36
1.5.2.2 Symmetry and Antisymmetry
1.5.2.2.1 Symmetric tensor
A second-order tensor A is symmetric, i.e.:
sym
A A , if the tensor is equal to its transpose:
symmetric is if

A A A = =
ji ij
components in T
A A (1.118)
in matrix form:
(
(
(

= =
33 23 13
23 22 12
13 12 11
A A A
A A A
A A A
sym T
A A A A (1.119)
From the above it is clear that a symmetric second-order tensor has 6 independent
components, namely:
11
A ,
22
A ,
33
A ,
12
A ,
23
A ,
13
A .
According to equation (1.118), a symmetric tensor can be represented by:
) (
2
1
) (
2
1
2
T
ji ij ij
ji ij ij
ji ij ij ij
ji ij
A A A + = + =
+ =
+ = +
=
A A A
A A A
A A A A
A A
(1.120)
A fourth-order tensor C , whose components are
ijkl
C , may have the following types of
symmetries:
Minor symmetry:
jilk ijlk jikl ijkl
C C C C = = =
(1.121)
Major symmetry:
klij ijkl
C C =
(1.122)
A fourth-order tensor that does not exhibit any kind of symmetry has 81 independent
components. If the tensor C has only minor symmetry, i.e. symmetry in ) 6 ( ji ij = , and
symmetry in ) 6 ( lk kl = , the tensor features 36 independent components. If besides
presenting minor symmetry it also provides major symmetry, the tensor features 21
independent components.
1.5.2.2.2 Antisymmetric tensor
A tensor A is antisymmetric (also called skew-symmetric tensor or skew tensor), i.e.:
skew
A A :
ric antisymmet is if A A A = =
ji ij
components in T
A A (1.123)
which broken down into its components is the same as:
(
(
(


= =
0
0
0
23 13
23 12
13 12
A A
A A
A A
skew T
A A A
(1.124)
Therefore, an antisymmetric second-order tensor has 3 independent components, namely:
12
A ,
23
A ,
13
A .
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
37
Under the conditions expressed in (1.123), an antisymmetric tensor can be represented by:
) (
2
1
) (
2
1
2
T
ji ij ij
ji ij ij
ji ij ij ij
A A A = =
=
= +
A A A
A A A
A A A A

(1.125)
Let us consider an antisymmetric second-order tensor denoted by W, then satisfy the
above relationship (1.125):
) (
2
1
) (
2
1
) (
2
1
il jk jl ik kl il jk kl jl ik kl ji ij ij
= = = W W W W W W (1.126)
Using the relation between the Kronecker delta and the permutation symbol given by
(1.62), i.e.
lkr ijr il jk jl ik
= , the equation (1.126) is rewritten as:
lkr ijr kl ij
W W
2
1
= (1.127)
Expanding the term
lkr kl
W , for the dummy indices ( k , l ), we can obtain the following
nonzero terms:
r r r r r r lkr kl 23 32 13 31 32 23 12 21 31 13 21 12
W W W W W W W + + + + + = (1.128)
thus,
r lkr kl
lkr kl
lkr kl
lkr kl
w
w r
w r
w r
2
2 2 3
2 2 2
2 2 1
3 12 21 12
2 13 31 13
1 23 32 23
=

= = + = =
= = = =
= = + = =

W
W W W W
W W W W
W W W W
(1.129)
In which we assume the following variables have changed:
(
(
(

=
(
(
(


=
(
(
(

=
0
0
0
0
0
0
0
0
0
1 2
1 3
2 3
23 13
23 12
13 12
32 31
23 21
13 12
w w
w w
w w
ij
W W
W W
W W
W W
W W
W W
W
(1.130)
Hence, we introduce the axial vector w
r
associated with the antisymmetric tensor, W, as:
3 3 2 2 1 1

e e e w w w + + = w
r
(1.131)
The magnitude of the axial vector w
r
is given by:
2
12
2
13
2
23
2
3
2
2
2
1
2
2
W W W + + = + + = = = w w w w w w
r r r
(1.132)
Substituting (1.129) into (1.127) and by considering that
rij ijr
= we obtain:
rij r ij
w = W (1.133)
Multiplying both sides of the equation (1.133) by
kij
we can obtain:
k rk r kij rij r ij kij
w w w 2 2 = = = W (1.134)
where we have applied the relation
rk kij rij
2 = , which was evaluated in Problem 1.7,
thus we can conclude that:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
38
ij kij k
w W
2
1
= (1.135)
Graphical representation of the antisymmetric tensor components and its corresponding
axial vector, in the Cartesian system, is shown in Figure 1.15.












Figure 1.15: Antisymmetric tensor components and the axial vector.
Let a
r
and b
r
be arbitrary vectors and W be an antisymmetric tensor, it follows that:
b W a b W a a W b
r
r
r
r r
r
= =
T

(1.136)
when b a
r
r
= , it holds that:
0 ) ( = = a a W a W a
r r r r
: (1.137)
NOTE: Note that ) ( a a
r r
is a symmetric second-order tensor. Later on we will show that
the result of the double contraction between a symmetric tensor and an antisymmetric
tensor equals zero.
Let us consider an antisymmetric tensor W and an arbitrary vector a
r
. The components of
the scalar product a W
r
are given by:
3 33 2 32 1 31
3 23 2 22 1 21
3 13 2 12 1 11
3 3 2 2 1 1
3
2
1
a W a W a W
a W a W a W
a W a W a W
a W a W a W a W
+ + =
+ + =
+ + =
+ + =
i
i
i
i i i j ij
(1.138)
Bearing in mind that the normal components are equal to zero for an antisymmetric tensor,
i.e., 0
11
= W , 0
22
= W , 0
33
= W , the scalar product (1.138) becomes:
( )

+ =
+ =
+ =

2 32 1 31
3 23 1 21
3 13 2 12
3
2
1
a W a W
a W a W
a W a W
i
i
i
i
a W
r
(1.139)
The above components are the same as the result of the algebraic operation a
r
r
w :

23 1
W = w

1
x

2
x

3
x

12
W

12
W

23
W

13
W

13
W

13 2
W = w

12 3
W = w

3 3 2 2 1 1

e e e w w w + + = w
r


23
W

(
(
(


=
0
0
0
23 13
23 12
13 12
W W
W W
W W
W
ij

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
39
( ) ( ) ( )
( ) ( ) ( )
3 2 32 1 31 2 3 23 1 21 1 3 13 2 12
3 2 1 1 2 2 3 1 1 3 1 3 2 2 3
3 2 1
3 2 1
3 2 1



e e e
e e e
e e e
a
a W a W a W a W a W a W
a a a a a a
a a a
+ + + + + =
+ + + + =
=
w w w w w w
w w w
r r
w

(1.140)
where
32 23 1
W W = = w ,
31 13 2
W W = = w ,
21 12 3
W W = = w . Then, given an antisymmetric
tensor W and the axial vector w
r
associated with W, it holds that:
a a W
r r r
= w (1.141)
for any vector a
r
. The property (1.141) could easily have been obtained by taking into
account the components of W given by (1.133), i.e.:
( ) ( )
i k j ijk k jik j k ik i
w w a a W
r
r
r
= = = = w a a a W (1.142)
The vector w
r
can be represented by its magnitude, = w
r
, and by the unit vector
codirectional with w
r
, i.e.
*
1

e = w
r
. Then, the equation (1.141) can still be expressed as:
a e a a W
r r r r
= =
*
1

w (1.143)
Additionally, we can choose two unit vectors
*
2

e ,
*
3

e , which make up an orthonormal basis


with the unit vector
*
1

e , (see Figure 1.16), so that:


*
2
*
1
*
3
*
1
*
3
*
2
*
3
*
2
*
1

;

;

e e e e e e e e e = = = (1.144)






Figure 1.16: Orthonormal basis defined by the axial vector.
By representing the vector a
r
in this new basis,
*
3
*
3
*
2
*
2
*
1
*
1

e e e a a a a + + =
r
, the relationship
shown in (1.143) obtains the form below:
[ ] a e e e e
e e e e e e e e
e e e e a e a W
e e
0
r
43 42 1 43 42 1 43 42 1
r r

=
= + + =
+ + = =
= =
=
)

(
)

( )

(
)

(

*
3
*
2
*
2
*
3
*
2
*
3
*
3
*
2
*
2
*
3
*
1
*
3
*
3
*
2
*
1
*
2
*
1
*
1
*
1
*
3
*
3
*
2
*
2
*
1
*
1
*
1
*
1



a a a a a
a a a

(1.145)
Thus, an antisymmetric tensor can be represented, in the space defined by the axial vector,
as follows:
)

(
*
3
*
2
*
2
*
3
e e e e W = (1.146)
Note that by using the antisymmetric tensor representation shown in (1.146), the
projections of the tensor W according to directions
*
1

e ,
*
2

e and
*
3

e are respectively:

*
3

e

*
1

e
3

e

2

e

1

e

*
2

e

*
1

e = w
r

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
40
*
2
*
3
*
3
*
2
*
1

;

;

e e W e e W 0 e W = = =
r
(1.147)
We can also verify that:
[ ]
[ ]

= =
= =


*
3
*
3
*
2
*
2
*
3
*
2
*
3
*
2
*
2
*
3
*
2
*
2
*
3
*
3
*
2
*
3

)

(

)

(

e e e e e e e W e
e e e e e e e W e
(1.148)
Then, the tensor components of W in the basis formed by the orthonormal basis
*
1

e ,
*
2

e ,
*
3

e , are given by:


(
(
(

=
0 0
0 0
0 0 0
*

ij
W (1.149)
In Figure 1.17 we can see these components and the axial vector representation. Note that
if we take any basis that is formed just by rotation along the
*
1

e -axis, the components of


W in this new basis will be the same as those provided in (1.149).










Figure 1.17: Antisymmetric tensor components in the space defined by the axial vector.
1.5.2.2.3 Additive decomposition. Symmetric and antisymmetric part
Any arbitrary second-order tensor A can be split additively into a symmetric and an
antisymmetric part:
skew sym T T
skew sym
A A A A A A A
A A
+ = + + =
43 42 1 43 42 1
) (
2
1
) (
2
1

(1.150)
or, into its components:
) (
2
1
) (
2
1
ji ij
skew
ij ji ij
sym
ij
and A A A A A A = + = (1.151)
If A and B are arbitrary second-order tensors, it holds that:
( ) ( ) ( ) [ ]
[ ] A B A A B B A
A B A A B A A B A A B A A B A


= + =
+ =
(

+ =
sym T T T
T T T
T
T T
sym
T
2
1
2
1
2
1
(1.152)

*
3

e

*
1

e


*
2

e

1
x

2
x
3
x


*
1

e = w
r


(
(
(

=
0 0
0 0
0 0 0
*

ij
W
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
41
Problem 1.15: Show that 0 = W : is always true when is a symmetric second-order
tensor and W is an antisymmetric second-order tensor.
Solution:
(scalar) )

( )

(
ij ij jk il lk ij k l lk j i ij
W W W = = = e e e e W : :
Thus,
43 42 1 43 42 1 3 2 1
33 33
32 32
31 31
3 3
23 23
22 22
21 21
2 2
13 13
12 12
11 11
1 1
W
W
W
W
W
W
W
W
W
W
W
W W

=
j j j j j j ij ij

Taking into account the characteristics of a symmetric and an antisymmetric tensor, i.e.
21 12
= ,
13 31
= ,
23 32
= , and 0
33 22 11
= = = W W W ,
12 21
W W = ,
13 31
W W = ,
23 32
W W = , the equation above becomes:
0 = W :
Problem 1.16: Show that a) M Q M M Q M
r r r r
=
sym
; b)
skew skew sym sym
B A B A B A : : : + =

where M
r
is a vector, and Q, A , B are arbitrary second-order tensors.
Solution:
a) ( ) M Q M M Q M M Q Q M M Q M
r r r r r r r r
+ = + =
skew sym skew sym

Since the relation ( ) 0

= =
43 42 1
r r r r
tensor symmetric
skew skew
M M Q M Q M : holds, it follows that:
M Q M M Q M
r r r r
=
sym

b)
skew skew sym sym
skew skew sym skew skew sym sym sym
skew sym skew sym
B A B A
B A B A B A B A
B B A A B A
: :
: : : :
: :
+ =
+ + + =
+ + =
= =
43 42 1 43 42 1
0 0
) ( ) (

Then, it is also valid that:
skew skew skew sym sym sym
B A B A B A B A : : : : = = ;
Problem 1.17: Let T be an arbitrary second-order tensor, and n
r
be a vector. Check if the
relationship n T T n
r r
= is valid.

Solution:
l l l l
l kl k
l ik kl i
l k kl i i
e
e
e
e e e T n

) (


)

(

3 3 2 2 1 1
T n T n T n
T n
T n
T n
+ + =
=
=
=

r
and
l l l l
l lk k
l ki lk i
i i k l lk
e
e
e
e e e n T

) (



)

(
3 3 2 2 1 1
T n T n T n
T n
T n
n T
+ + =
=
=
=

r

With the above we can prove that
lk k kl k
T n T n , then:
n T T n
r r

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
42
If T is a symmetric tensor, it follows that the relationship n T T n
r r
=
sym sym
holds.
Problem 1.18: Obtain the axial vector w
r
associated with the antisymmetric tensor
skew
) ( a x
r r
.
Solution: Let z
r
be an arbitrary vector, it then holds that:
z w z a x
r r r r r
=
skew
) (
where w
r
is the axial vector associated with
skew
) ( a x
r r
. Using the definition of an
antisymmetric tensor:
[ ] [ ] x a a x a x a x a x
r r r r r r r r r r
= =
2
1
) ( ) (
2
1
) (
T skew

and by replacing it with z w z a x
r r r r r
=
skew
) ( , we obtain:
[ ] [ ] z w z x a a x z w z x a a x
r r r r r r r r r r r r r r
= = 2
2
1

By using the equation [ ] ) ( a x z z x a a x
r r r r r r r r
= , (see Eq. (1.105)), the above equation
becomes:
[ ] z w z x a a x z z x a a x
r r r r r r r r r r r r r
= = = 2 ) ( ) (
with the above we can conclude that:
skew
with associated vector axial the is ) ( ) (
2
1
a x x a w
r r r r r
=
1.5.2.3 Cofactor Tensor. Adjugate of a Tensor
Let A be a second-order tensor and a
r
, b
r
be arbitrary vectors then there is then a unique
tensor ) cof(A , known as the cofactor of A , as we can see below:
) ( ) ( ) ( b A a A b a A
r
r
r
r
= ) cof( (1.153)
We can also define the adjugate of A as:
[ ]
T
) (A A cof ) adj( =

(1.154)
which satisfies the following condition:
[ ] ) adj( ) adj(
T T
A A = (1.155)
The components of ) cof(A are obtained by expressing the equation (1.153) in terms of its
components, i.e.:
[ ] [ ]
kr jp ijk tpr it r kr p jp ijk r p tpr it
A A ) cof( b A a A b a ) cof( = = A A
(1.156)
By multiplying both sides of the equation by
qpr
and by also considering that
tq qpr tpr
2 = , we can conclude that:
[ ] [ ]
[ ]
kr jp qpr ijk iq
kr jp qpr ijk
tq
qpr tpr it kr jp ijk tpr it
A A ) cof(
A A ) cof( A A ) cof(


2
1
2
=
= =
=
A
A A
43 42 1

(1.157)
1.5.2.4 Tensor Trace
Lets start by defining the trace of the basis )

(
j i
e e :
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
43
ij j i j i
= = e e e e

)

( Tr (1.158)
Thus, we can define the trace of a second-order tensor A as follows:
33 22 11
)

( )

( )

( ) (
A A A
A A A Tr A A Tr Tr
+ + =
= = = = =
ii ij ij j i ij j i ij j i ij
e e e e e e A
(1.159)
And, the trace of the dyad ) ( v u
r r
can be evaluated as:
v u
e e e e v u v u
r r
r r r r


= + + =
= = = = =
3 3 2 2 1 1
)

( )

( ) ( ) (
v u v u v u
v u v u v u Tr v u Tr Tr
i i ij j i j i j i j i j i

(1.160)
NOTE: As we will show later, the tensor trace is an invariant, i.e. it is independent of the
coordinate system.
Let A , B be arbitrary tensors, then:
The transposed tensor trace is equal to the tensor trace:
( ) ( ) A A Tr Tr =
T
(1.161)
The trace of ( ) B A + is the sum of traces:
( ) ( ) ( ) B A B A Tr Tr Tr + = + (1.162)
Proving this is very simple:
( ) ( ) ( )
( ) ( ) ( ) [ ] ( ) ( )
33 22 11 33 22 11 33 33 22 22 11 11
B B B A A A B A B A B A
Tr Tr Tr
+ + + + + = + + + + +
+ = + B A B A

(1.163)
The scalar product trace ) ( B A becomes:
( ) [ ]
[ ]
( ) A B B A
e e
e e e e B A


= = =
=
=

)

(
)

( )

(
Tr B A
Tr B A
B A Tr Tr
li il
im
m i jl lm ij
m l lm j i ij
4 4 3 4 4 2 1


(1.164)
and, the double scalar product ( : ) can be expressed in trace terms as:
( ) ( )
) ( ) (
) (
) (
B A B A
B A B A
B A
B A
B A


= =
= =
= =
= =
=

T T
kk
T
kk
T
kl
kl
T
il ik kl
kl
T
lj kj
jl jk il ik il ik lj kj
ij ij
Tr Tr
B A B A
B A B A
B A


3 2 1
3 2 1
:
(1.165)
Similarly, it is possible to show that:
( ) ( ) ( )
ki jk ij
C B A Tr Tr Tr = = = B A C A C B C B A
(1.166)
ii
A Tr = ) (A
[ ]
jj ii
A A Tr Tr Tr = = ) ( ) ( ) (
2
A A A
( ) ( )
li il
A A Tr Tr = =
2
A A A ; ( ) ( )
ki jk ij
A A A Tr Tr = =
3
A A A A
(1.167)

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
44
Problem 1.19: Let T be a second-order tensor. Show that:
( ) ( ) ( ) ( )
m
m
T
m
T
T
m
and T T T T Tr Tr = = .
Solution:
( ) ( ) ( )
m
T T T T T
T
m
T T T T T T T T = = = L L
For the second demonstration we can use the trace property ( ) ( ) T T Tr Tr =
T
, thus:
( ) ( ) ( )
m
T
m
m
T
T T T Tr Tr Tr = =
1.5.2.5 Particular Tensors
1.5.2.5.1 Unit Tensors
The second-order unit tensor, also called the identity tensor, is defined as:
j i i i j i ij
e e e e e e 1

= = = 1 (1.168)
where 1 is the matrix with the components of tensor 1.
ij
is the Kronecker delta symbol
defined in (1.48).
Fourth-order unit tensors can be defined as follows:
l l l l
e e e e e e e e 1 1

= = =
k j i ijk k j i j ik
I I (1.169)
l l l l
e e e e e e e e 1 1

= = =
k j i ijk k j i jk i
I I (1.170)
l l l l
e e e e e e e e 1 1

= = =
k j i ijk k j i k ij
I I (1.171)
Taking into account the fourth-order unit tensors defined above, it holds that:
( ) ( ) ( )
( ) ( )
A
e e e e
e e e e e e e e A
=
= =
= =
j i ij j i k j ik
j i q kp pq j ik q p pq k j i j ik


A A
A A
l l
l l l l

: : I
(1.172)
and
( ) ( ) ( )
( ) ( )
T
j i ji j i k jk i
j i q kp pq jk i q p pq k j i jk i
A
e e e e
e e e e e e e e A
=
= =
= =


A A
A A
l l
l l l l

: : I
(1.173)
and
( ) ( ) ( )
( ) ( )
1 A
e e e e
e e e e e e e e A
) (


Tr
A A
A A
=
= =
= =
j i ij kk j i k k ij
j i q kp pq k ij q p pq k j i k ij


l l
l l l l
: : I
(1.174)
The symmetric part of the fourth-order unit tensor I is defined as:
( ) ( )
jk i j ik ijk
components in sym

l l l
+ = + =
2
1
2
1
I 1 1 1 1 I I (1.175)
The property of the tensor product is presented below. Consider a second-order unit
tensor,
j i ij
e e 1

= . Then, the tensor product can be defined as:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
45

( ) ( ) ( )
l l l l
e e e e e e e e 1 1

= =
j k i k ij k k j i ij

(1.176)
which is the same as:
( )
l l
e e e e 1 1

= =
k j i j ik
I (1.177)
and the tensor product is defined as:
( ) ( ) ( )
j k i k ij k k j i ij
e e e e e e e e 1 1

= =
l l l l
(1.178)
or
( )
l l
e e e e 1 1

= =
k j i jk i
I (1.179)
The antisymmetric part of I is defined as:
( ) ( )
jk i j ik
skew
ijk
components in skew

l l l
= =
2
1
2
1
I 1 1 1 1 I (1.180)
With a second-order tensor A and a vector b
r
, the following relationships are valid:
b 1 b
r r
=
sym sym
A A A A = = : : I I ;
ii
A Tr = = ) (A 1 A:
( ) ( )
li il
A A Tr Tr = = = A A A 1 A
2 2
:
( ) ( )
ki jk ij
A A A Tr Tr = = = A A A A 1 A
3 3
:
(1.181)
Problem 1.20: Show that ) (T 1 T Tr = : , where T is an arbitrary second-order tensor.
Solution:
) (

T
e e e e 1 T
Tr
T T T
T
T
=
= = =
=
=
jj ii ij ij
jl ik kl ij
l k kl j i ij


: :

1.5.2.5.2 Levi-Civita Pseudo-Tensor
The Levi-Civita Pseudo-Tensor, also known as the Permutation Tensor, is a third-order pseudo-
tensor and is denoted by:
k j i ijk
e e e

= (1.182)
which is not a true tensor in the strict meaning of the word, and whose components
ijk

were defined in (1.55), the permutation symbol.
1.5.2.6 Determinant of a Tensor
The determinant of a tensor is a scalar and is expressed as:
4 4 3 4 4 2 1
T
k j i ijk k j i ijk
A
A A
3 2 1 3 2 1
) ( A A A A A A det = =
(1.183)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
46
It is also an invariant (independent of the adopted system). Demonstrating the equation
above (1.183) can be done starting directly from the determinant:
3 2 1
3 2 31 3 3 2 21 2 3 2 11 1
3 2 3 31 3 2 2 21 3 2 1 11
22 13 23 12 31 32 13 33 12 21 32 23 33 22 11
33 32 31
23 22 21
13 12 11
) ( ) ( ) (
) ( ) ( ) (
) (
k j i ijk
k j jk k j jk k j jk
k j jk k j jk k j jk
A A A
A A A A A A A A A
A A A A A A A A A
A A A A A A A A A A A A A A A
A A A
A A A
A A A
det



=
+ + =
+ =
+ =
= = A A
(1.184)
Some determinant properties with second-order tensors are described below:
1 ) ( = 1 det
(1.185)
We can conclude from (1.183) that:
) ( ) ( A A det det =
T

(1.186)
We can also show that:
) ( ) ( ) ( B A B A det det det = , ) ( ) (
3
A A det det = where is a scalar (1.187)
A tensor ) (A is said to be singular if 0 ) ( = A det .
If you exchange two rows or columns, the determinant sign changes.
If all elements of a row or column equal zero, the determinant is also zero.
If you multiply all the elements of a row or column by a constant c (scalar), the
determinant is A c .
If two or more rows (or column) are linearly dependent, the determinant is zero.
Problem 1.21: Show that
kq jp rt rjk tpq
A A A = A .
Solution:
We start with the following definition:
3 2 1 3 2 1 k j r tpq rjk tpq k j r rjk
A A A A A A = = A A
(1.188)
and also taking into account that the term
tpq rjk
can be replaced by (1.61):
rp jt kq rt kp jq kt jp rq kp jt rq kt jq rp kq jp rt
kq kp kt
jq jp jt
rq rp rt
tpq rjk




+ + =
=
(1.189)
Then, by substituting (1.189) into (1.188) we can obtain:
( ) ( ) ( )
qk pj tr rjk kq jp rt rjk
qk pj jk t qk pj jk t qk pj jk t
q t p p q t t p q p t q t q p q p t tpq
A A A A A A
A A A A A A A A A
A A A A A A A A A A A A A A A A A A

= =
+ + =
+ + =
3 3 2 2 1 1
3 2 1 3 2 1 3 2 1 3 2 1 3 2 1 3 2 1
A

Problem 1.22: Show that
kq jp rt tpq rjk
A A A
6
1
= A .
Solution:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
47
Starting with the definition
kq jp rt rjk tpq
A A A = A , (see Problem 1.21), and by multiplying
both sides of the equation by
tpq
, we obtain:
kq jp rt tpq rjk tpq tpq
A A A = A
(1.190)
Using the property defined in expression (1.62), we obtain
6 = = =
tt pp tt tp tp pp tt tpq tpq
. Then, the relationship (1.190) becomes:
kq jp rt tpq rjk
A A A
6
1
= A
Problem 1.23: Let a
r
, b
r
be arbitrary vectors and be a scalar. Show that:
( ) b a b a 1
r
r
r
r
+ = +
2 3
det
(1.191)
Solution: The determinant of A is given by
3 2 1 k j i ijk
A A A = A . If we denote by
j i ij ij
b a A + , thus,
1 1 1
b a A
i i i
+ = ,
3 3 3
b a A
k k k
+ = ,
3 3 3
b a A
k k k
+ = , then
the equation in (1.191) can be rewritten as:
( ) ( )( )( )
3 3 2 2 1 1
b a b a b a det
k k j j i i ijk
+ + + = + b a 1
r
r
(1.192)
By developing the equation (1.192), we obtain:
( ) [
]
3 2 1
3
3 2 1
2
2 3 1
2
1 3 2
2
3 2 1
2
3 1 2
2
2 1 3
2
3 2 1
3
b b b a a a b b a a b b a a b a b a
b a b a b a det
k j i k j i j k i i k j
k j i k i j j i k k j i ijk


+ + + +
+ + + + = + b a 1
r
r

Note that:
3
123
3
3 2 1
3
= =
k j i ijk
,
0
0
) ( ) ( ) (
) (
2 1 1 2 213 2 1 2 1 123 2 1 3 3 2 1
3 1 1 3 3 1 3 1 3 1 2 2 3 1
2
1 1 2 2 3 3
2
1 23 2 3 1 3 12
2
3 2 1 3 1 2 2 1 3
2

= = =
= = =
= + + = + +
= + +

b b a a b b a a b b a a b b a a
b b a a b b a a b b a a b b a a
b a b a b a b a b a b a
b a b a b a




j i ij k j i ijk
k i k i j k i ijk
i i j j k k
k j i ijk k i j ijk j i k ijk



b a
r
r

0
3 2 1
= b b b a a a
k j i ijk

Notice that, there was no need to expand the terms
2 3 1 j k i ijk
b b a a ,
3 2 1 k j i ijk
b b a a , and
3 2 1
b b b a a a
k j i ijk
to realize that these terms equal zero, since
0 ) (
2 3 1 2 3 1
= =
j j j k i ijk
b b b b a a a a
r r
, similarly for other terms.
Taking into account the above considerations we can prove that:
( ) b a b a 1
r
r
r
r
+ = +
2 3
det
For the particular case when 1 = the above equation becomes:
( ) b a b a 1
r
r
r
r
+ = + 1 det
Then, it is simple to prove that ( ) 0 = b a
r
r
det , since
( ) [ ] 0 ) (
3 2 1
3
3 2 1
3
= = = a a a b a
r r r
r
r
b b b b b b a a a det
k j i ijk


The following relations are satisfied:
[ ] [ ] ) )( ( ) ( ) ( ) ( 1 ) ( ) (
2
b b a a b a b a b a a b b a 1
r r
r r
r
r
r
r
r
r r
r r
r
+ + + = + + det
(1.193)
where , are scalars. If 0 = we can regain the equation ( ) b a b a 1
r
r
r
r
+ = + 1 det ,
(see Problem 1.23). If = we obtain:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
48
( ) ( ) ( ) ( ) ( )( )
( ) ( )
(

+ =
(

+ + + = + +


2
2
2
2 1
1
b a b a
b b a a b a b a b a a b b a 1
r
r
r
r
r r
r r
r
r
r
r
r
r r
r r
r

det

(1.194)
where we have used the property ( ) ( )( ) ( )
2 2
b a b b a a b a
r
r
r r
r r
r
r
= , (see Problem 1.1).
It is also true that:
( ) [ ] [ ] ) ( ) ( ) ( ) (
3 2 2 3
B B A A B A B A det adj Tr adj Tr det det + + + = + (1.195)
Moreover, in the particular case when 1 = , 1 A = , b a B
r
r
= , and bearing in mind that
( ) 0 = b a
r
r
det , and ( ) 0 b a =
r
r
cof , we can conclude that:
( ) ( ) [ ] [ ] b a 1 b a 1 b a 1
r
r
r
r
r
r
+ = + = + = + 1 1
j i
b a Tr Tr det det (1.196)
which has already been demonstrated in Problem 1.23.
We next show that the following property is valid:
[ ] [ ] ) ( ) ( ) ( ) ( ) ( c b a A c A b A a A
r
r
r r
r
r
= det
(1.197)
To achieve this goal we start with the definition of the scalar triple product given in (1.69),
i.e. ( )
k j i ijk
c b a = c b a
r
r
r
, and by multiplying both sides of this equation by the
determinant of A we obtain:
( ) A A c b a
k j i ijk
c b a =
r
r
r
(1.198)
It was proven in Problem 1.21 that
rk qj pi pqr ijk
A A A = A , thus:
( )
[ ] [ ] ) ( ) ( ) ( ) ( ) ( ) (
) )( )( (
a A c A b A c A b A a A
A A c b a
r r
r
r
r
r
r
r
r

= =
= = =
k rk j qj i pi pqr k j i rk qj pi pqr k j i ijk
c A b A a A c b a A A A c b a

(1.199)
1.5.2.7 Inverse of a Tensor
We use the notation
1
A to denote the inverse of A , which is defined as follows:
if 1 A A A A A A = =


1 1 1
0 (1.200)
Or in indicial notation:
if
ij kj ik kj ik ij
= =

A A A A A
1 1 1
0 A (1.201)
To obtain the inverse of a tensor we start from the definition of the adjugate tensor given
in (1.153), i.e. ) ( ) ( ) ( b A a A b a A
r
r
r
r
= ) adj(
T
. Then by applying the dot product
between an arbitrary vector d
r
and this equation we obtain:
[ ] { } [ ] [ ]
[ ]
3 2 1
r r
r
r r
r
r r
r
r r
r
r
c
d A A b A a A
d 1 b A a A d b A a A d b a A
=

=
= =
1
) ( ) (
) ( ) ( ) ( ) ( ) (
T
) adj(
(1.202)
In equation (1.199) we demonstrated that ( ) [ ] ) ( ) ( ) ( c A b A a A A b a c
r
r
r
r
r r
= thus,
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
49
[ ] { } [ ] ( ) d A b a A d A A b A a A d b a A
r r
r
r r
r
r r
r


= =
1 1
) ( ) ( ) (
T
) adj( (1.203)
Denoted by ) ( b a p
r
r r
= , the above equation (1.203) can be rearranged as follows:
[ ] { }
[ ]
[ ] [ ] [ ] d b a A A d b a A
A A
A A
r r
r
r r
r
=
=
=

) ( ) (
1
1
1
: : ) adj(
d p A d p ) adj(
d A p d p ) adj(
i k ki i k ki
i ki k i k ki
(1.204)
Thus, we can conclude that:
[ ] =
1
A A A) adj( [ ] [ ]
T
) cof( ) adj( A
A
A
A
A
1 1
1
= =


(1.205)
Let A and B be invertible tensors, the following properties hold:

( )
( )
[ ]
1 1
1 1
1
1
1 1 1
) ( ) (
1
) (


=
=
=
=
A A
A A
A A
A B B A
det det

(1.206)
The following notation will be used to represent the inverse transpose:
1 1
) ( ) (


T T T
A A A (1.207)
Next, we prove the relation ) adj( ) adj( ) adj( A B B A = holds. To do this, we take the
definition of the inverse of a tensor given in (1.205) as a starting point:
[ ] [ ]
[ ] [ ]
( ) [ ] [ ]
[ ]
[ ] [ ]
[ ] [ ] ) adj( ) adj( ) adj(
) adj( ) adj(
) adj(
) adj( ) adj(
) adj( ) adj(
) adj( ) adj(
A B B A
A B
B A
B A
B A A B B A B A
A B A B B A
A
A
B
B
A B




=
= =
= =


1
1 1 1 1

(1.208)
where we have used the property B A B A = . Similarly, it is possible to show that
[ ] [ ] ) cof( ) cof( ) cof( B A B A = .
Procedure for obtaining the inverse of the matrix A
1) Obtain the cofactor matrix: ) (A cof as follows:
Consider the matrix A as:
(
(
(

=
33 32 31
23 22 21
13 12 11
A A A
A A A
A A A
A
(1.209)
We define the matrix M, where the component
ij
M is the determinant of the resulting
matrix by removing the
th
i row and the
th
j column, i.e.:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
50
(
(
(
(
(
(
(
(

=
22 21
12 11
23 21
13 11
23 22
13 12
32 31
12 11
33 31
13 11
33 32
13 12
32 31
22 21
33 31
23 21
33 32
23 22
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
A A
M (1.210)
Thus, we define the cofactor matrix as:
ij
j i
M A
+
= ) 1 ( ) ( cof (1.211)
2) Obtain the adjugate matrix, ) adj(A , as follows:
[ ]
T
) (A A cof ) adj( = (1.212)
3) Obtain the inverse matrix:
A
A
A
) adj(
=
1

(1.213)
So, the relation [ ] 1 A A A = ) adj( holds, where 1 is the identity matrix.
Taking into account (1.64), we can express the components of the first, second, and third
row of the cofactor matrix, (1.211), as:
k j ijk i 3 2 1
A A M = ,
k j ijk i 3 1 2
A A M = ,
k j ijk i 2 1 3
A A M = ,
respectively.
Problem 1.24: Let A be an arbitrary second-order tensor. Show that there is a nonzero
vector 0 n
r
r
so that 0 n A
r
r
= if and only if 0 ) ( = A det , Chadwick (1976).
Solution: Firstly, we show that, if 0 n A A
r
r
= 0 ) ( det . Secondly, we show that, if
0 ) ( = A A 0 n det
r
r
.
We assume that 0 ) ( = A A det , and we choose an arbitrary basis } , , { h g f
r
r
r
(linearly
independent). We apply these vectors in the definition seen in (1.197):
( ) [ ] ) ( ) ( ) ( h A g A f A A h g f
r
r
r r
r
r
=
Due to the fact that 0 ) ( = A A det , the implication is that:
[ ] 0 ) ( ) ( ) ( = h A g A f A
r
r
r

Thus, we can conclude that the vectors ) ( f A
r
, ) ( g A
r
, ) ( h A
r
, are linearly dependent.
This implies that there are nonzero scalars , , so that:
( ) 0 n A 0 h g f A 0 h A g A f A
r
r
r r
r
r r r
r
r
= = + + = + + ) ( ) ( ) (
where 0 h g f n
r r
r
r
r
+ + = since } , , { h g f
r
r
r
is linearly independent, (see Problem 1.10).
Now we choose two vectors k
r
, m
r
, which are linearly independent to n
r
. We apply
definition (1.199) once more:
( ) [ ] ) ( ) ( ) ( n A m A k A A n m k
r r
r
r r
r
=
Considering that 0 n A
r
r
= , and ( ) 0 n m k
r r
r
owing to the fact that k
r
, m
r
, n
r
are linearly
independent, we can conclude that:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
51
( ) 0 0
0
= =

A A n m k
43 42 1
r r
r

1.5.2.8 Orthogonal Tensors
Orthogonal tensors play an important role in continuum mechanics. A second-order tensor
( Q) is said to be orthogonal when the transpose (
T
Q ) is equal to the inverse (
1
Q ), i.e.
1
= Q Q
T
. Then, it follows that:
ij kj ki jk ik
T T
= = = = Q Q Q Q ; 1 Q Q Q Q (1.214)
A proper orthogonal tensor has the following properties:
The inverse of Q is equal to the transpose (orthogonality):
T
Q Q =
1
(1.215)
The tensor Q is a proper, rotation tensor, if:
1 ) ( + = Q Q det (1.216)
If 1 = Q , the orthogonal tensor is said to be improper (a reflection tensor).
If A and B are orthogonal tensors, the resulting tensor C B A = is also an orthogonal
tensor, i.e.:
T T T T
C B A A B A B B A C = = = = =

) ( ) (
1 1 1 1
(1.217)
Consider two arbitrary vectors a
r
and b
r
. An orthogonal transformation applied to these
vectors becomes:
b Q b a Q a
r
r
r
r
= =
~
;
~

(1.218)
And the dot product between these new vectors ( a
r
~
) and ( b
r
~
) is given by:
k k j
kj
ij ik k j ij k ik i i
T
b a b Q Q a b Q a Q b a = = =
= = =
=
3 2 1
r
r
r
3 2 1
r
r
r
r
r

) )( (
~
~
) ( ) (
~
~
b a b Q Q a b Q a Q b a
1
(1.219)
So, it is also true when b a
r
r
~
~
= , thus
2
2
~ ~ ~
a a a a a a
r r r
r r r
= = = . Therefore, we can conclude
that in an orthogonal transformation, the magnitude vectors and the angle between them
are maintained, (see Figure 1.18).





NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
52







Figure 1.18: Orthogonal transformation applied to vectors.
1.5.2.9 Positive Definite Tensor, Negative Definite Tensor and Semi-
Definite Tensors
A tensor is said to be positive definite when the following notations hold:
Tensorial notation Indicial notation Matrix notation
0

> x T x

0 >
j ij i
x T x

0 > x x T
T

(1.220)
for all vectors 0 x
r

. Conversely, a tensor is said to be negative definite when these notations


hold:
Tensorial notation Indicial notation Matrix notation
0

< x T x

0 <
j ij i
x T x

0 < x x T
T

(1.221)
for all vectors 0 x
r
r
.
A tensor is said to be semi-positive definite if 0

x T x for all vectors 0 x
r

. Similarly, we
define a semi-negative definite tensor when the following holds: 0

x T x .
If
j i ij
x x T = = = )

(

x x T x T x : , then the derivative of with respect to x
r
is given
by:
( )
i ik ki i ik j kj jk i ij j ik ij
k
j
i ij j
k
i
ij
k
x T T x T x T x T x T
x
x
x T x
x
x
T
x
+ = + = + =

(1.222)
Thus, we can conclude that:
sym sym
T
x x
x T
x
2

2
=


=



(1.223)
Remember that it is also true that x T x x T x

=
sym
, therefore if the symmetric part of
a tensor is positive definite the tensor is too.
NOTE: As we will demonstrate later, the eigenvalues must be positive for T to be
positive definite. The proof is in the subsection Spectral Representation of Tensors.
Problem 1.25: Let F be an arbitrary second-order tensor. Show that the resulting tensors
F F C =
T
and
T
F F b = are symmetric tensors and semi-positive definite tensors. Also check in
what condition are C and b positive definite tensors.
Solution: Symmetry:
Q
a
r

b
r



a
r
~

b
r
~


b a b a
b b
a a
r
r
r
r
r
r
r
r
=
=
=
~
~
~
~

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
53
b F F F F F F b
C F F F F F F C
= = = =
= = = =


T T T T T T T
T T T T T T T
) ( ) (
) ( ) (

Thus, we have shown that F F C =
T
and
T
F F b = are symmetric tensors.
To prove that the tensors F F C =
T
and
T
F F b = are semi-positive definite tensors, we
start with the definition of a semi-positive definite tensor, i.e., a tensor A is semi-positive
definite if 0

x A x holds, for all 0 x
r

. Thus:

0

( )

( )

( )

(

) (

) (

2 2
= =
= =
= =



x x
x x x x
x x x x x x x x
T
T T
T T T
F F
F F F F
F F F F F F F F

Or in indicial notation:
0 0
) )( ( ) )( (
) ( ) (
2 2
= =
= =
= =
i ik i ki
j jk i ik j kj i ki
j jk ik i j ij i j kj ki i j ij i
F F
F F F F
F F b F F C
x x
x x x x
x x x x x x x x

Thus, we proved that F F C =
T
and
T
F F b = are semi-positive definite tensors. Note
that
2

x x x = F C equals zero, when 0 x
r

, if 0 x
r
=

F . Furthermore, by definition
0 x
r
=

F with 0 x
r

if and only if 0 ) ( = F det , (see Problem 1.24). Then, the tensors


F F C =
T
and
T
F F b = are positive definite if and only if 0 ) ( F det .
1.5.2.10 Additive Decomposition of Tensors
Given two arbitrary tensors S , 0 T , and a scalar , we can represent S by means of the
following additive decomposition of tensors:
T S U U T S = + = where (1.224)
Note that, depending on the value of , we have an infinite number of possibilities for
representing S . But, if 0 ) ( ) ( = =
T T
T U U T Tr Tr , the additive decomposition is unique.
From the relationship in (1.224), we can evaluate the value of as follows:
) ( ) ( ) ( ) (
0
T T T T T T T
T T T U T T T S T U T T T S = + = + =
=
Tr Tr Tr Tr
43 42 1

(1.225)
) (
) (
T
T
T T
T S

=
Tr
Tr
(1.226)
For example, let us suppose that 1 T = . In this case is evaluated as follows:
3
) (
) (
) (
) (
) (
) (
) ( S
1
S
1 1
1 S
T T
T S Tr
Tr
Tr
Tr
Tr
Tr
Tr
= = = =

T
T
(1.227)
We can then define U as:
dev
S 1
S
S T S U = =
3
) ( Tr
(1.228)
Thus:
dev sph dev
S S S 1
S
S + = + =
3
) ( Tr
(1.229)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
54
NOTE: 1
S
S
3
) ( Tr
=
sph
is the spherical part of the tensor S , and 1
S
S S
3
) ( Tr
=
dev
is
known as a deviatoric tensor.
Now suppose that T is given by ) (
2
1
T
S S T + = then can be evaluated as follows:
[ ]
[ ]
1
) ( ) (
4
1
) (
2
1
) (
) (
=
+ +
+
= =

T T T
T T
T
T
S S S S
S S S
T T
T S
Tr
Tr
Tr
Tr
(1.230)
We can then define U as ) (
2
1
) (
2
1
T T
S S S S S T S U = + = = . Then we obtain S
represented by the additive decomposition as follows:
skew sym T T
S S S S S S S + = + + = ) (
2
1
) (
2
1
(1.231)
which is the same as the equation obtained in (1.150) in which we split the tensor into
symmetric and antisymmetric parts.
Problem 1.26: Find a fourth-order tensor P so that
dev
A A = : P , where A is a second-
order tensor.
Solution: Taking into account the additive decomposition into spherical and deviatoric parts,
we obtain:
1
A
A A A 1
A
A A A
3
) (
3
) ( Tr Tr
= + = + =
dev dev dev sph

Referring to the definition of fourth-order unit tensors seen in (1.172), and (1.174), where
the relations 1 A A ) ( Tr = : I and A A = : I hold, we can now state:
A 1 1 A A A 1
A
A A : : : : |
.
|

\
|
= |
.
|

\
|
= = =
3
1
3
1
3
1
3
) (
I I I I I
Tr
dev

Therefore, we can conclude that:
1 1 =
3
1
I P
The tensor P is known as a fourth-order projection tensor, Holzapfel(2000).
1.5.3 Transformation Law of the Tensor Components
The tensor components depend on the coordinate system, so, if the coordinate system is
changed due to a rotation so do the tensor components. The tensor components between
these coordinate systems are interrelated to each other by the component transformation
law, which is defined below, (see Figure 1.19).
Consider a Cartesian coordinate system ( )
3 2 1
, , x x x formed by the orthogonal basis
( )
3 2 1

,

e e e , (see Figure 1.20). In this system, an arbitrary vector v


r
is represented by its
components as follows:
3 3 2 2 1 1

e e e e v v v v v + + = =
i i
r
(1.232)

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
55




















Figure 1.19: Transformation law of the tensor components.














Figure 1.20: Rotation of the Cartesian system.

2
x

1
x

3
x

1
x

2
x

3
x

1


1


1


2

e

3

e

1

e

3

e

1

e

2

e
COMPONENTS
Representation of a tensor in
a coordinate system
COORDINATE SYSTEM
I

COORDINATE SYSTEM
II
COMPONENT
TRANSFORMATION
LAW
TENSORS
Mathematical interpretation of physical concepts
(Independent of the coordinate system)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
56
The components,
i
v , are represented in matrix form as:
(
(
(

= = =
3
2
1
) (
v
v
v
v v
i i
v
r
(1.233)
Now consider a new coordinate system ( )
3 2 1
, , x x x represented by the orthogonal basis
( )
3 2 1

,

e e e , (see Figure 1.20). In this new system, the vector v


r
is represented by
j j
e

v . As
mentioned before, a tensor is independent of the adopted system, so:
j j k k
e e v

v v = =
r
(1.234)
To obtain the components of a tensor in a given system one need only make the dot
product between the tensor and the system basis:
i i
i j j ki k
i j j i k k
e e e e
e e
e e e e
+ + =
=
=

)

(

(

3 3 2 2 1 1
v v v v
v v
v v
(1.235)
Or in matrix form:
(
(
(

+ +
+ +
+ +
=
(
(
(

3 3 3 2 2 1 1
2 3 3 2 2 1 1
1 3 3 2 2 1 1
3
2
1

)

(

)

(

)

(
e e e e
e e e e
e e e e
v v v
v v v
v v v
v
v
v
(1.236)
After restructuring, the previous equation looks like:
j i i j ij
a e e e e
e e e e e e
e e e e e e
e e e e e e




3
2
1
3 3 3 2 3 1
2 3 2 2 2 1
1 3 1 2 1 1
3
2
1




= =
(
(
(

(
(
(




=
(
(
(

v
v
v
v
v
v
(1.237)
Or in indicial notation:
j ij i
a v v =
(1.238)
where we have introduced the transformation matrix
ij
a A as:
(
(
(




=
(
(
(

3 3 2 3 1 3
3 2 2 2 1 2
3 1 2 1 1 1
3 3 3 2 3 1
2 3 2 2 2 1
1 3 1 2 1 1






e e e e e e
e e e e e e
e e e e e e
e e e e e e
e e e e e e
e e e e e e
ij
a A
(
(
(

=
33 32 31
23 22 21
13 12 11
a a a
a a a
a a a
a
ij
A
(1.239)
The matrix ( A) is not symmetric, i.e.
T
A A . With reference to the scalar product
( ) ( )
j i j i j i j i
x x x x , cos , cos

= = e e e e , (see equation (1.4)), the relationship in (1.237) is
expressed by means of the direction cosines as:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
57
{
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )
{
v
v
(
(
(

(
(
(




=
(
(
(

3
2
1
3 3 2 3 1 3
3 2 2 2 1 2
3 1 2 1 1 1
3
2
1

, cos , cos , cos
, cos , cos , cos
, cos , cos , cos
v
v
v
v
v
v
4 4 4 4 4 4 4 3 4 4 4 4 4 4 4 2 1
A
x x x x x x
x x x x x x
x x x x x x

v v A =

(1.240)
The direction cosines of a vector are those of the angles between the vector and the three
coordinate axes. According to Figure 1.20 we can verify that ( )
1 1 1
, cos cos x x = ,
( )
2 1 1
, cos cos x x = and ( )
3 1 1
, cos cos x x = .
In the equation (1.235) we have projected the vector onto
i
e

. Now, we can project the


vector onto
i
e

:
v v =
=
=
=
T
ji j i
ji j ki k
i j j i k k
a
a
A
v v
v v
v v

e e e e

(1.241)
Therefore, it is also true that:
j ji i
a e e =

(1.242)
The inverse relationship of equation (1.240) is obtained as follows:
v v v v = =
1 1 1
A A A A (1.243)
and by comparing the equations (1.243) with (1.241) we can conclude that the matrix A is
an orthogonal matrix, i.e.:
ij kj ki
T T
a a = = =


notation
Indicial
1
1 A A A A
(1.244)
Second-order tensor
Consider a coordinate system formed by the orthogonal basis
i
e

then, how the basis


changes from the
i
e

system to a new one represented by the orthogonal basis


i
e

. This is
illustrated in transformation law as
i ik k
a e e =

, which allow us to represent a second-order
tensor T as follows:
j i ij
j i jl ik kl
j jl i ik kl
l k kl
a a
a a
e e
e e
e e
e e T
=
=
=
=




T
T
T
T
(1.245)
Then, the transformation law of the components between systems for a second-order
tensor is given by:
T
jl kl ik jl ik kl ij
a a a a A T A T
form Matrix
= = = T T T (1.246)
Third-order tensor
A third-order tensor ( S ) can be shown in two systems represented by orthogonal bases
i
e


and
i
e

as follows:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
58
k j i ijk
k j i kn jm il lmn
k kn j jm i il lmn
n m l lmn
a a a
a a a
e e e
e e e
e e e
e e e S
=
=
=
=




S
S
S
S
(1.247)
In conclusion the components of the third-order tensor in the new basis (
i
e

) are:
kn jm il lmn ijk
a a a S S =
(1.248)
The following table summarizes the transformation law of the components according to
the tensor rank:
rank
from ( ) ( )
3 2 1 3 2 1
, , , , x x x x x x
to
from ( ) ( )
3 2 1 3 2 1
, , , , x x x x x x
to

0 (scalar) = =
1 (vector)
j ij i
S a S =
j ji i
S a S =
2
kl jl ik ij
S a a S =
kl lj ki ij
S a a S =
3
lmn kn jm il ijk
S a a a S =
lmn nk mj li ijk
S a a a S =
4
mnpq lq kp jn im ijkl
S a a a a S =
mnpq ql pk nj mi ijkl
S a a a a S =
(1.249)

Problem 1.27: Obtain the components of T , given by the transformation:
T
A T A T =
where the components of T and A are shown, respectively, as
ij
T and
ij
a . Afterwards,
given that
ij
a are the components of the transformation matrix, represent graphically the
components of the tensors T and T on both systems.
Solution: The expression
T
A T A T = in symbolic notation is given by:
)

(
)

(
)

( )

( )

( )

(
k r kq pq rp
k r ql sp kl pq rs
k l kl q p pq s r rs b a ab
a a
a a
a a
e e
e e
e e e e e e e e
=
=
=
T
T
T T

To obtain the components of T one only need make the double scalar product with the
basis )

(
j i
e e , the result of which is:
jq pq ip ij
kj ri kq pq rp bj ai ab
j i k r kq pq rp j i b a ab
a a
a a
a a
T T
T T
T T
=
=
=

)

( )

( )

( )

( e e e e e e e e : :

The above eqaution is shown in matrix notation as:
T inverse T
= = A T A T A T A T
1

Since A is an orthogonal matrix, it holds that
1
= A A
T
. Thus, A T A T =
T
. The
graphical representation of the tensor components in both systems can be seen in Figure
1.21.




1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
59




















Figure 1.21: Transformation law of the second-order tensor components.
Problem 1.28: Let T be a symmetric second-order tensor and
T
I ,
T
I I ,
T
I I I be scalars,
where:
{ } ) ( ; ) (
2
1
; ) (
2 2
T T T
T T T T
det Tr T Tr = = = = I I I I I I I
ii

Show that
T
I ,
T
I I ,
T
I I I are invariant with a change of basis.
Solution:
a) Taking into account the transformation law for the second-order tensor components
given in (1.249), i.e.
kl jl ik ij
a a T T = or in matrix form
T
A T A T = . Then,
ii
T is:
T
I a a
kk kl kl kl il ik ii
= = = = T T T T
Hence we have proved that
T
I is independent of the adopted system.
b) To prove that
T
I I is an invariant, one only need show that ) (
2
T Tr is one also, since
2
T
I
is already an invariant.
) ( ) (
) )( ( ) ( ) (
2
2
T T T T T
T T T T T
Tr Tr
T T
T T
T T T T Tr Tr
= = =
=
=
= = = =

:
:
pl pl
pq kl
lq
jq jl
kp
ip ik
pq jq ip kl jl ik ij ij
a a a a
a a a a
3 2 1 3 2 1


c)
) ( ) ( ) ( ) ( ) ( ) ( ) (
1 1
T A T A A T A T det det det det det det det = = = =
= =

43 42 1 43 42 1
T T
T
Consider now four sets of coordinate systems, represented by ( )
3 2 1
, , x x x , ( )
3 2 1
, , x x x ,
( )
3 2 1
, , x x x and ( )
3 2 1
, , x x x , (see Figure 1.22), and consider also the following
transformation matrices:
A: Transformation matrix from ( )
3 2 1
, , x x x to ( )
3 2 1
, , x x x ;

1
x
2
x

3
x

11
T

21
T

31
T

12
T

32
T

22
T

33
T

13
T

23
T

3
x

2
x

1
x

11
T

21
T

31
T

12
T

32
T

22
T

33
T

13
T

23
T

T
A T A T =
A T A T =
T


3
x

1
x

2
x
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
60
B : Transformation matrix from ( )
3 2 1
, , x x x to ( )
3 2 1
, , x x x ;
C : Transformation matrix from ( )
3 2 1
, , x x x to ( )
3 2 1
, , x x x .











Figure 1.22: Transformations matrices between several systems.
If we consider a v column matrix made up of components of v
r
in the coordinate system
( )
3 2 1
, , x x x , the components of this vector in the system ( )
3 2 1
, , x x x are given by:
v v A = (1.250)
and the inverse transformation of relation (1.250) is:
v v =
T
A
(1.251)
Now, starting with the system ( )
3 2 1
, , x x x , the components of the vector in the system
( )
3 2 1
, , x x x are given by:
v v = B (1.252)
and the inverse transformation is:
v v =
T
B
(1.253)
By substituting the equation (1.250) into (1.252) we obtain:
v v BA = (1.254)
The resulting matrix BA is also an orthogonal matrix, and shows the transformation
matrix from ( )
3 2 1
, , x x x to ( )
3 2 1
, , x x x , (see Figure 1.22). The inverse form of (1.254) is
evaluated by substituting (1.253) into (1.251), the result of which is:
v v =
T T
B A
(1.255)
This equation could have been obtained by using equation (1.254), i.e.:
( ) ( ) ( ) ( ) v v v v v v = = = =
T T
B A B A BA BA BA BA
1 1 1 1 1
(1.256)
Then, it is easy to find the components of the vector in the coordinate system ( )
3 2 1
, , x x x ,
(see Figure 1.22):
v v v v = =
T T T
C B A CBA
form inverse
(1.257)

X
X X
B

T
B B =
1

A

T
A A =
1


T T T
) (BA B A =
BA
X
T T T T
C B A CBA = ) (
CBA

T
C C =
1

C
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
61
1.5.3.1 Component Transformation Law in Two Dimensions (2D)
Now, consider two sets of coordinate systems, shown in Figure 1.23.







Figure 1.23: Transformation of a coordinate system in 2D.
The transformation matrix from ( y x ) to ( y x ) is given by direction cosines, (see
Figure 1.23), as:
(
(
(

=
(
(
(

=


1 0 0
0 ) cos( ) cos(
0 ) cos( ) cos(
1 0 0
0
0
22 21
12 11
y y x y
y x x x
a a
a a


A
(1.258)
By using trigonometric identities we can deduce that:
) cos( ) cos( ) cos( = = =
y y x x y y x x
, ) sin(
2
cos ) cos( = |
.
|

\
|
=


y x
,
) sin(
2
cos ) cos( = |
.
|

\
|
+ =


x y

(1.259)
Thus, the transformation matrix in 2D is dependent on a single parameter, , i.e.:
(

=
) cos( ) sin(
) sin( ) cos(


A (1.260)
Another way to prove (1.260) is by considering the vector position of the point P in both
systems, (see Figure 1.24).
Moreover, in view of Figure 1.24, said coordinates are interrelated as shown below:

|
.
|

\
|

+ =
+ =


2
cos ) cos(
) cos( ) cos(
P P P
P P P
y x y
y x x

+ |
.
|

\
|

=
|
.
|

\
|

+ =
) cos(
2
cos
2
cos ) cos(


P P P
P P P
y x y
y x x

+ =
+ =

) cos( ) sin(
) sin( ) cos(


P P P
P P P
y x y
y x x

(1.261)
Or in matrix form:
(

=
(

=
(


P
P
P
P tion transforma
Inverse
P
P
P
P
y
x
y
x
y
x
y
x

) cos( ) sin(
) sin( ) cos(

) cos( ) sin(
) sin( ) cos(
1





(1.262)
Since
T
A A =
1
, it is true that:

y y


y x

=
x x

x
y
x
y

x y





+

=
=

2
2
x y
y x
y y

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
62
(


=
(

P
P
P
P
y
x
y
x
) cos( ) sin(
) sin( ) cos(


(1.263)













Figure 1.24: Transformation of a coordinate system in 2D.





Problem 1.29: Find the transformation matrix between the systems: z y x , , and z y x , , .
These systems are represented in Figure 1.25.















Figure 1.25: Rotation.

y y =
x
y
z z =
x
z z =
y
x
x






P
x
P
y

x
y
x
y

P
y
r
r

P

P
x

)
c
o
s
(

P
x
)
c
o
s
(

P
x



)
s
i
n
(

P
y
)
s
i
n
(

P
y


)
s
i
n
(

P
x
)
s
i
n
(

P
x


)
c
o
s
(

P
y
)
c
o
s
(

P
y


P
y

P
y


1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
63
Solution: The coordinate system z y x , , can be obtained by different combinations of
rotations as follows:

Rotation along the z -axis











Rotation along the y -axis















Rotation along the z -axis

















from z y x , , to z y x , ,

(
(
(

=
1 0 0
0 cos sin
0 sin cos


C
with 360 0
y y =
x
y
z z =
x
z z =
y
x
x





from z y x , , to z y x , ,

(
(
(

=
1 0 0
0 cos sin
0 sin cos


A
with 360 0
x
y
z z =
x
y


from z y x , , to z y x , ,

(
(
(


=


cos 0 sin
0 1 0
sin 0 cos
B
with 180 0
y y =
x
y
z z =
x
z
x




x
z z =
z
x
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
64
The transformation matrix from ( z y x , , ) to ( z y x , , ), (see Figure 1.22), is given by:
CBA D =
After multiplying the matrices, we obtain:
(
(
(

+
+
=



cos sin sin sin cos
sin sin ) cos cos sin cos sin ( ) cos sin sin cos cos (
cos sin ) sin cos cos cos (sin ) sin sin cos cos (cos
D
The angles , , are known as Euler angles and were introduced by Leonhard Euler to
describe the orientation of a rigid body motion.

Problem 1.30: Let T be a second-order tensor whose components in the Cartesian system
( )
3 2 1
, , x x x are given by:
( )
(
(
(


= = =
1 0 0
0 3 1
0 1 3
T
ij ij
T T
Given that the transformation matrix between two systems, ( )
3 2 1
, , x x x - ( )
3 2 1
, , x x x , is:
(
(
(
(
(
(

=
0
2
2
2
2
0
2
2
2
2
1 0 0
A
Obtain the tensor components
ij
T in the new coordinate system ( )
3 2 1
, , x x x .
Solution: As defined in equation (1.249), the transformation law for second-order tensor
components is:
kl jl ik ij
a a T T =
To enable the previous calculation to be carried out in matrix form we use:
[ ] [ ] [ ]
T
j l l k k i ij
a a

T T =
Thus
T
A T A T =
(
(
(
(
(
(
(

(
(
(


(
(
(
(
(
(

=
0 0 1
2
2
2
2
0
2
2
2
2
0
1 0 0
0 3 1
0 1 3
0
2
2
2
2
0
2
2
2
2
1 0 0
T
On carrying out the operation of the previous matrices we now have:
(
(
(

=
4 0 0
0 2 0
0 0 1
T
NOTE: As we can verify in the above example, the components of the tensor T , in the
new basis, have one particular feature, i.e. the off-diagonal terms are equal to zero. The
question now is: Given an arbitrary tensor T , is there a transformation which results in the
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
65
off-diagonal terms being zero? This type of problem is called the eigenvalue and eigenvector
problem.
1.5.4 Eigenvalue and Eigenvector Problem
As we have seen, the scalar product between a second-order tensor T and a vector (or unit
vector n

) leads to a vector. In other words, projecting a second-order tensor onto a


certain direction results in a vector that does not necessarily have the same direction as n

,
(see Figure 1.26(a)).
The aim of the eigenvalue and eigenvector problem is to find a direction n

, in such a way
that the resulting vector, n T t
n

)

(
=
r
, coincides with it, (see Figure 1.26 (b)).














Figure 1.26: Projecting a tensor onto a direction.
Let T be a second-order tensor. A vector n

is said to be eigenvector of T if there is a scalar


, called the eigenvalue, so that:
n n T

= (1.264)
The equation (1.264) can be rearranged in indicial notation as:
( ) ( ) 0 n 1 T
r
= =
=
=




notation
Tensorial
i j ij ij
i i j ij
i j ij
0 n T
0 n n T
n n T


(1.265)
The previous set of homogeneous equations only have nontrivial solution, i.e. 0 n
r

, if and
only if:
0 ; 0 ) ( = =
ij ij
T det 1 T (1.266)
The determinant (1.266) is called the characteristic determinant of the tensor T , explicitly given
by:
n T t
n
=

(
r

n



1
x

2
x

3
x
n n T t
n

)

(
= =
r

n

- principal direction of T
- eigenvalue of T associated
with the direction n

.
b) Principal direction. a) Projection of T onto an
arbitrary plane.
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
66
0
33 32 31
23 22 21
13 12 11
=



T T T
T T T
T T T

(1.267)
Developing this determinant, we obtain the characteristic polynomial, which is shown by a
cubic equation in :
0
2 3
= +
T T T
I I I I I I (1.268)
where
T
I ,
T
I I ,
T
I I I are the principal invariants of T , and are defined in components terms
as:
ii
I T Tr = = ) (T
T

[ ]
( ) ( ) [ ] { }
( ) [ ] { }
{ }
{ } [ ] ) cof( Tr M T T T T
T T T T
Tr T T T T
T T Tr T Tr T Tr
Tr Tr
T
e e
e e e e e e e e
T T
T
= = =
=
=
=
=

ii ji ij kk ii
il jk kl ij kk ii
l i jk kl ij kl kl ij ij
l k kl j i ij l k kl j i ij
I I
2
1
2
1

2
1

)

( )

(
2
1
) ( ) (
2
1
2 2


3 2 1
) (
k j i ijk ij
I I I T T T T det = = = T
T

(1.269)
where
ii
M is the matrix trace defined in equation (1.210),
33 22 11
M M M M + + =
ii
. More
explicitly the invariants are given by:
( ) ( ) ( )
22 31 32 21 13 23 31 33 21 12 23 32 33 22 11
21 12 22 11 31 13 33 11 32 23 33 22
22 21
12 11
33 31
13 11
33 32
23 22
33 22 11
T T T T T T T T T T T T T T T
T T T T T T T T T T T T
T T
T T
T T
T T
T T
T T
T T T
+ =
+ + =
+ + =
+ + =
T
T
T
I I I
I I
I

(1.270)
If T is a symmetric tensor, the principal invariants are summarized as follows:
22
2
13 11
2
23 33
2
12 23 12 13 23 13 12 33 22 11
2
23
2
13
2
12 33 22 33 11 22 11
33 22 11
T T T T T T T T T T T T T T T
T T T T T T T T T
T T T
+ + =
+ + =
+ + =
T
T
T
I I I
I I
I

(1.271)
The eigenvalues,
3 2 1
, , , are found by solving the characteristic polynomial (1.268).
Once the eigenvalues are evaluated, the eigenvectors are found by applying equation
(1.265), i.e.
i j ij ij
0 n T =
) 1 (
1

) ( ,
i j ij ij
0 n T =
) 2 (
2

) ( ,
i j ij ij
0 n T =
) 3 (
3

) ( . These
eigenvectors constitute a new space denoted as the principal space.
If T is a symmetric tensor, the principal space is defined by an orthonormal basis and all
eigenvalues are real numbers. If the three eigenvalues are different,
3 2 1
, the three
principal directions are unique. If two of them are equal, e.g.
3 2 1
= , we can state that
the principal direction,
) 3 (

n , associated with the eigenvalue


3
, is unique, and, any
direction defined in the plane normal to
) 3 (

n is a principal direction, and othorgonality is



jk

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
67
the only constraint to determining
) 1 (

n and
) 2 (

n . If
3 2 1
= = , any direction is principal.
A tensor that has three equal eigenvalues is called a Spherical Tensor, (see Appendix A-The
Tensor ellipsoid).
The T -components in the principal space are only made up of normal components, i.e.:
(
(
(

=
(
(
(

=
3
2
1
2
2
1
0 0
0 0
0 0
0 0
0 0
0 0
T
T
T
T
ij

(1.272)
Therefore, the principal invariants can also be evaluated by:
3 2 1
T T T + + =
T
I ,
3 1 3 2 2 1
T T T T T T + + =
T
I I ,
3 2 1
T T T =
T
I I I (1.273)
whose values must match the values obtained in (1.270), since they are invariant with a
change of basis.
If T is a spherical tensor, i.e. T T T T = = =
3 2 1
, it holds that
T T
I I I 3
2
= ,
3
T =
T
I I I .
Let W be an antisymmetric tensor. The principal invariants of W are given by:
[ ]
0
0
0
0
0
0
0
2
) (
) ( ) (
2
1
0 ) (
2
12 12 13 13 23 23
12
12
13
13
23
23
2
2 2
=
=
+ + =

= =
= =
W
W
W
W
W W
W
I I I
I I
I

W W W W W W
W
W
W
W
W
W
Tr
Tr Tr
Tr
(1.274)
where
2
12
2
13
2
23
2
2
W W W + + = = = w w w
r r r
as defined in (1.132). Then, the characteristic
equation for an antisymmetric tensor is reduced to:
( ) 0 0 0
2 2 2 3 2 3
= + = + = +
W W W
I I I I I I (1.275)
In this case, one eigenvalue is real and equal to zero and the others are imaginary roots:
i 1 0 0
) 2 , 1 (
2 2 2 2
= = = = = + (1.276)
1.5.4.1 The Orthogonality of the Eigenvectors
Consider a symmetric second-order tensor T . By the definition of eigenvalues, given in
(1.264), if
1
,
2
,
3
are the eigenvalues of T , then it follows that:
) 3 (
3
) 3 ( ) 2 (
2
) 2 ( ) 1 (
1
) 1 (

;

n n T n n T n n T = = = (1.277)
Applying the dot product between
) 2 (

n and
) 1 (
1
) 1 (

n n T = , and the dot product between
) 1 (

n and
) 2 (
2
) 2 (

n n T = we obtain:
) 2 ( ) 1 (
2
) 2 ( ) 1 (
) 1 ( ) 2 (
1
) 1 ( ) 2 (


n n n T n
n n n T n


=
=
(1.278)
Since T is symmetric, it holds that
) 2 ( ) 1 ( ) 1 ( ) 2 (

n T n n T n = , so:
) 1 ( ) 2 (
2
) 2 ( ) 1 (
2
) 1 ( ) 2 (
1

n n n n n n = = (1.279)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
68
( ) 0

) 2 ( ) 1 (
2 1
= n n (1.280)
To satisfy the equation (1.280), with 0
2 1
, the following must be true:
0

) 2 ( ) 1 (
= n n (1.281)
Similarly, it is possible to show that 0

) 3 ( ) 1 (
= n n and 0

) 3 ( ) 2 (
= n n and then we can
conclude that the eigenvectors are mutually orthogonal, and constitute an orthogonal basis,
(see Figure 1.27), where the transformation matrix between systems is:
(
(
(

=
(
(
(

=
) 3 (
3
) 3 (
2
) 3 (
1
) 2 (
3
) 2 (
2
) 2 (
1
) 1 (
3
) 1 (
2
) 1 (
1
) 3 (
) 2 (
) 1 (


n n n
n n n
n n n
n
n
n
A (1.282)
















Figure 1.27: Diagonalization.
Problem 1.31: Show that the following relations are invariants:
4
3
4
2
4
1
3
3
3
2
3
1
2
3
2
2
2
1
; ; C C C C C C C C C + + + + + +
where
1
C ,
2
C ,
3
C are the eigenvalues of the second-order tensor C .
Solution: Any combination of invariants is also an invariant, so, on this basis, we can try to
express the above expressions in terms of their principal invariants.
( ) ( )
C C
C
C
I I I C C C C C C C C C C C C C C C I
I I
2 2
2 2
3
2
2
2
1 3 2 3 1 2 1
2
3
2
2
2
1
2
3 2 1
2
= + + + + + + + = + + =
4 4 4 4 3 4 4 4 4 2 1

So, we have proved that
2
3
2
2
2
1
C C C + + is an invariant. Similarly, we can obtain:
2
C C C
2
C C C
C C C C
I I I I I I I I I I C C C
I I I I I I I C C C
2 4 4
3 3
4 4
3
4
2
4
1
3 3
3
3
2
3
1
+ + = + +
+ = + +


1
x

2
x

3
x

1
x

2
x
diagonalization

3
x

1
x

2
x

3
x

22
T

12
T
12
T

11
T

23
T

23
T

13
T

13
T

33
T

2
T

1
T

3
T

) 1 (

n

) 2 (

n
) 3 (

n
Principal Space

T
A T A T =
A T A T =
T

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
69

Problem 1.32: Let Q be a proper orthogonal tensor, and E be an arbitrary second-order
tensor. Show that the eigenvalues of E do not change with the following orthogonal
transformation:
T
Q E Q E =
*

Solution: We can prove this as follows:
( )
( )
( )
( ) [ ]
( ) ( ) ( )
( )
( )
( )
( )
( ) [ ]
( ) ( ) ( )
( )
kp kp
jp kp kp ik
jp kp kp ik
kp jp ik jp kp ik
ij jp kp ik
ij ij
T
T
T T
T

=
=
=
=
=
=
=
=
=
=
=
=



E det
Q det E det Q det
Q E Q det
Q Q Q E Q det
Q E Q det
E det
det
det det det
det
det
det
det

0

0
*
1 1
*
1 E
Q 1 E Q
Q 1 E Q
Q 1 Q Q E Q
1 Q E Q
1 E
43 42 1 3 2 1

Thus, we have proved that E and
*
E have the same eigenvalues.
1.5.4.2 Solution of the Cubic Equation
Let T be a symmetric second-order tensor. The roots of the characteristic equation
( 0
2 3
= +
T T T
I I I I I I ) are all real numbers, and are expressed as:
3 3
4
3
cos 2
3 3
2
3
cos 2
3 3
cos 2
3
2
1
T
T
T
I
S
I
S
I
S
+
(

|
.
|

\
|
+ =
+
(

|
.
|

\
|
+ =
+
(

|
.
|

\
|
=

(1.283)
where
|
.
|

\
|
= = = =

=
T
Q R
T
I
I I I
I I I
Q
R
S
I I I
R
2
arccos ;
27
;
27
2
3
;
3
;
3
3
3 3 2

T
T
T T T T

where is in radians.
(1.284)
By restructuring the solution (1.283) in matrix form, we obtain:
4 4 4 4 4 4 4 4 3 4 4 4 4 4 4 4 4 2 1
4 43 4 42 1
part Deviatoric
part Spherical
3
2
1
3
4
3
cos 0 0
0
3
2
3
cos 0
0 0
3
cos
2
1 0 0
0 1 0
0 0 1
3
0 0
0 0
0 0
(
(
(
(
(
(
(

|
.
|

\
|
+
|
.
|

\
|
+
|
.
|

\
|
+
(
(
(

=
(
(
(

S
I
T

(1.285)
where we clearly distinguish the spherical and the deviatoric part of the tensor in the
principal space. Note that, if T is a spherical tensor the following relationship holds
T T
I I I 3
2
= , then 0 = S .
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
70

Problem 1.33: Find the principal values and directions of the second-order tensor T ,
where the Cartesian components of T are:
( )
(
(
(


= = =
1 0 0
0 3 1
0 1 3
T
ij ij
T T

Solution: We need to find nontrivial solutions for ( )
i j ij ij
0

= n T , which are constrained


by 1

=
j j
n n (unit vector). As we have seen, the nontrivial solution requires that:
0 =
ij ij
T
Explicitly, the above equation is:
0
1 0 0
0 3 1
0 1 3
33 32 31
23 22 21
13 12 11
=



=



T T T
T T T
T T T

Developing the above determinant, we can obtain the cubic equation:
[ ]
0 8 14 7
0 1 ) 3 ( ) 1 (
2 3
2
= +
=

We could have obtained the characteristic equation directly in terms of invariants:
7 ) (
33 22 11
= + + = = = T T T T T Tr
ii ij
I
T

( ) 14
2
1
22 21
12 11
33 31
13 11
33 32
23 22
= + + = =
T T
T T
T T
T T
T T
T T
T T T T
ij ij jj ii
I I
T

8
3 2 1
= = =
k j i ijk ij
I I I T T T T
T

Then, using the equation in (1.268), the characteristic equation is:
0 8 14 7 0
2 3 2 3
= + = +
T T T
I I I I I I
On solving the cubic equation we obtain three real roots, namely:
4 ; 2 ; 1
3 2 1
= = =
We can also verify that:
8
14 1 4 4 2 2 1
7 4 2 1
3 2 1
1 3 3 2 2 1
3 2 1
= =
= + + = + + =
= + + = + + =
T
T
T
I I I
I I
I

Thus, we can see that the invariants are the same as those evaluated previously.

Principal directions:
Each eigenvalue,
i
, is associated with a corresponding eigenvector,
) (

i
n . We can use the
equation in (1.265), i.e.
i j ij ij
0 n T =

) ( , to obtain the principal directions.
1
1
=
(
(
(

=
(
(
(

(
(
(



=
(
(
(

(
(
(




0
0
0
1 1 0 0
0 1 3 1
0 1 1 3
1 0 0
0 3 1
0 1 3
3
2
1
3
2
1
1
1
1
n
n
n
n
n
n

These become the following system of equations:



1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
71
1
0 0
0
0 2
0 2
2
3
2
2
2
1
3
2 1
2 1
2 1
= + + =

=
= =
)
`

= +
=
n n n n n
n
n n
n n
n n
i i

Then we can conclude that: [ ] 1 0 0

1
) 1 (
1
= =
i
n .
NOTE: This solution could have been directly determined by the specific features of the
T matrix. As the terms 0
32 31 23 13
= = = = T T T T imply that 1
33
= T is already a principal
value, then, consequently, the original direction is a principal direction.
2
2
=
(
(
(

=
(
(
(

(
(
(



=
(
(
(

(
(
(




0
0
0
2 1 0 0
0 2 3 1
0 1 2 3
1 0 0
0 3 1
0 1 3
3
2
1
3
2
1
2
2
2
n
n
n
n
n
n

=
= +
= =
0
0
0
3
2 1
2 1 2 1
n
n n
n n n n

The first two equations are linearly dependent, after which we need an additional equation:
2
1
1 2 1
1
2
1
2
3
2
2
2
1
= = = + + = n n n n n n n
i i

Thus:
(
(

= = 0
2
1
2
1

2
) 2 (
2 i
n
4
3
=
(
(
(

=
(
(
(

(
(
(



=
(
(
(

(
(
(




0
0
0
4 1 0 0
0 4 3 1
0 1 4 3
1 0 0
0 3 1
0 1 3
3
2
1
3
2
1
3
3
3
n
n
n
n
n
n

2
1
1 2 1
0 3
0
0
2
2
2
2
3
2
2
2
1
3
2 1
2 1
2 1
= = = + + =

=
=
)
`

=
=
n n n n n n n
n
n n
n n
n n
i i

Then:
(
(

= = 0
2
1
2
1

4
) 3 (
3
m
i
n
Afterwards, we summarize the eigenvalues and eigenvectors of T :
[ ]
(
(

= =
(
(

= =
= =
0
2
1
2
1

4
0
2
1
2
1

2
1 0 0

1
) 3 (
3
) 2 (
2
) 1 (
1
m
i
i
i
n
n
n

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
72
NOTE: The tensor components of this problem are the same as those used in Problem
1.30. Additionally, we can verify that the eigenvectors make up the transformation matrix,
A, between the original system, ( )
3 2 1
, , x x x , and the principal space, ( )
3 2 1
, , x x x , (see
Problem 1.30).
1.5.5 Spectral Representation of Tensors
Based on the solution of the equation in (1.268), if T is a symmetric second-order tensor
there are three real eigenvalues:
1
T ,
2
T ,
3
T each of which is associated with an
eigenvector, i.e.:
[ ]
[ ]
[ ]




) 3 (
3
) 3 (
2
) 3 (
1
) 3 (
3
) 2 (
3
) 2 (
2
) 2 (
1
) 2 (
2
) 1 (
3
) 1 (
2
) 1 (
1
) 1 (
1
n n n n T
n n n n T
n n n n T
=
=
=
i
i
i
(1.286)
The principal space is formed by the orthogonal basis
) 3 ( ) 2 ( ) 1 (

,

n n n , and the tensor


components are represented by their eigenvalues as:
(
(
(

= =
3
2
1
0 0
0 0
0 0
T
T
T
T T
ij

(1.287)
With reference to the fact that eigenvectors form a transformation matrix, A, so that:
T
A T A T = (1.288)
Since
T
A A =
1
, the inverse form is:
A T A T =
T

(1.289)
where
(
(
(

=
(
(
(

=
) 3 (
3
) 3 (
2
) 3 (
1
) 2 (
3
) 2 (
2
) 2 (
1
) 1 (
3
) 1 (
2
) 1 (
1
) 3 (
) 2 (
) 1 (


n n n
n n n
n n n
n
n
n
A (1.290)
Explicitly, the relation in (1.289) is given by:
A A A A A A
A A
(
(
(

(
(
(

(
(
(

(
(
(

=
(
(
(

(
(
(

(
(
(

=
(
(
(

+ + =
1 0 0
0 0 0
0 0 0
0 0 0
0 1 0
0 0 0
0 0 0
0 0 0
0 0 1
0 0
0 0
0 0




0 0
0 0
0 0




3 2 1
3
2
1
) 3 (
3
) 3 (
2
) 3 (
1
) 2 (
3
) 2 (
2
) 2 (
1
) 1 (
3
) 1 (
2
) 1 (
1
3
2
1
) 3 (
3
) 2 (
3
) 1 (
3
) 3 (
2
) 2 (
2
) 1 (
2
) 3 (
1
) 2 (
1
) 1 (
1
33 23 13
23 22 12
13 12 11
T T T
T
T T T
T
T
T
n n n
n n n
n n n
T
T
T
n n n
n n n
n n n
T T T
T T T
T T T
(1.291)
Whereas:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
73
) 3 ( ) 3 (
) 3 (
3
) 3 (
3
) 3 (
2
) 3 (
3
) 3 (
1
) 3 (
3
) 3 (
3
) 3 (
2
) 3 (
2
) 3 (
2
) 3 (
1
) 3 (
2
) 3 (
3
) 3 (
1
) 3 (
2
) 3 (
1
) 3 (
1
) 3 (
1
) 2 ( ) 2 (
) 2 (
3
) 2 (
3
) 2 (
2
) 2 (
3
) 2 (
1
) 2 (
3
) 2 (
3
) 2 (
2
) 2 (
2
) 2 (
2
) 2 (
1
) 2 (
2
) 2 (
3
) 2 (
1
) 2 (
2
) 2 (
1
) 2 (
1
) 2 (
1
) 1 ( ) 1 (
) 1 (
3
) 1 (
3
) 1 (
2
) 1 (
3
) 1 (
1
) 1 (
3
) 1 (
3
) 1 (
2
) 1 (
2
) 1 (
2
) 1 (
1
) 1 (
2
) 1 (
3
) 1 (
1
) 1 (
2
) 1 (
1
) 1 (
1
) 1 (
1




1 0 0
0 0 0
0 0 0




0 0 0
0 1 0
0 0 0




0 0 0
0 0 0
0 0 1
j i
T
j i
T
j i
T
n n
n n n n n n
n n n n n n
n n n n n n
n n
n n n n n n
n n n n n n
n n n n n n
n n
n n n n n n
n n n n n n
n n n n n n
=
(
(
(

(
(
(

=
(
(
(

(
(
(

=
(
(
(

(
(
(

=
=
=
A A
A A
A A

(1.292)
Then, it is possible to represent the components of a second-order tensor in function of
their eigenvalues and eigenvectors (spectral representation) as:
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1

j i j i j i ij
n n T n n T n n T T + + = (1.293)
As we can see, the tensor is represented as a linear combination of dyads and the above
representation in tensorial notation becomes:
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1

n n n n n n T + + = T T T (1.294)
or:

=
=
3
1
) ( ) (

a
a a
a
n n T T
Spectral representation of a
second-order tensor
(1.295)
which is the spectral representation of the tensor. Note that, in the above equation we have to
resort to the summation symbol, because the dummy index appears thrice in the
expression.
NOTE: The spectral representation in (1.295) could easily have been obtained from the
definition of the second-order unit tensor, given in (1.168), i.e.
i i
n n 1

= , which can also
be represented by means of the summation symbol as

=
=
3
1
) ( ) (

a
a a
n n 1 . Then, it follows
that:

= = =
= =
|
|
.
|

\
|
= =
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (

a
a a
a
a
a a
a
a a
n n n n T n n T 1 T T T (1.296)
where we have used the definition of eigenvalue and eigenvector
) ( ) (

a
a
a
n n T T = .
We now consider the orthogonal tensor R . The orthogonal transformation applied to the
unit vector N

leads to the unit vector n

, i.e.

N R n = . Therefore, it is also possible to


represent the orthogonal tensor R as follows:

= = =
= =
|
|
.
|

\
|
= =
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (


a
a a
a
a a
a
a a
N n N N R N N R 1 R R (1.297)
The spectral representation is very useful for making algebraic operations with tensors. For
example, tensor power in the principal space can be expressed as:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
74
( )
(
(
(

=
n
n
n
ij
n
3
2
1
0 0
0 0
0 0
T
T
T
T (1.298)
So, the spectral representation of
n
T is given by:

=
=
3
1
) ( ) (

a
a a n
a
n
n n T T (1.299)
Now, if we need the square root of the tensor, T , this can easily be obtained from the
spectral representation as:

=
=
3
1
) ( ) (

a
a a
a
n n T T (1.300)
Next, we can show that a positive definite tensor has positive eigenvalues. For this
purpose, we can consider a semi-positive definite tensor, T , by which the condition
0

x T x holds for all 0 x
r

. Replacing the tensor by its spectral representation, we


obtain:
0

0

0

3
1
) ( ) (
3
1
) ( ) (

|
|
.
|

\
|


=
=
x n n x
x n n x
x T x
a
a a
a
a
a a
a
T
T
(1.301)
Note that the result of the operation )

(
) (a
n x is a scalar, thus:
[ ]
0 )

( )

( )

(
0 )

( 0

2 ) 3 (
3
2 ) 2 (
2
2 ) 1 (
1
3
1
2 ) (
3
1
) ( ) (
0
+ +




= =
>
n x n x n x
n x x n n x
T T T
T T
a
a
a
a
a a
a
43 42 1
(1.302)
The above expression must hold for all 0 x
r

. If we take
) 1 (

n x = , the above equation is
reduced to 0 )

(
1
2 ) 1 ( ) 1 (
1
= T T n n . The same is true for
2
T and
3
T . Thus, we have
demonstrated that if a tensor is semi-positive definite, its eigenvalues are greater than or
equal to zero, i.e. 0
1
T , 0
2
T , 0
3
T . Therefore we can conclude that a tensor is positive
definite, i.e. 0

> x T x , if and only if its eigenvalues are positive and nonzero, i.e. 0
1
> T ,
0
2
> T , 0
3
> T . Consequently, the positive definite tensor trace is greater than zero. If the
positive definite tensor trace is zero, this implies that the tensor is the zero tensor.
The spectral representation of the fourth-order unit tensor, I , can be obtained starting
from the definition in (1.169), i.e.:

= =
= = =
3
1
3
1

a b
b a b a j i j i k j i j ik
e e e e e e e e e e e e
l l
I (1.303)
As I is an isotropic tensor, (see 1.5.8 Isotropic and Anisotropic tensors), then the
representation in (1.303) is also valid in any orthonormal basis,
) (

a
n , so:

= =
=
3
1
3
1
) ( ) ( ) ( ) (

a b
b a b a
n n n n I (1.304)
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
75
Similarly, we obtain the spectral representation for I and I as:
i j j i k j i jk i
e e e e e e e e

= =
l l
I
(1.305)

=
=
3
1 ,
) ( ) ( ) ( ) (

b a
a b b a
n n n n I (1.306)
and
k k i i k j i k ij
e e e e e e e e

= =
l l
I
(1.307)

=
=
3
1 ,
) ( ) ( ) ( ) (

b a
b b a a
n n n n I
(1.308)
Problem 1.34: Let w be an antisymmetric second-order tensor and V be a positive
definite symmetric tensor whose spectral representation is given by:

=
=
3
1
) ( ) (

a
a a
a
n n V
Show that the antisymmetric tensor w can be represented by:

=
=
3
1 ,
) ( ) (

b a
b a
b a
ab
n n w w
Demonstrate also that:

=
=
3
1 ,
) ( ) (

) (
b a
b a
b a
a b ab
n n V V w w w
Solution:
It is true that
( )
( )


=
= = =
=
= =
|
|
.
|

\
|
=
3
1 ,
) ( ) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (


b a
a a b
b
a
a a
a
a a
a
a a
w n n n
n n n n n n 1 w
r
w w w

where we have applied an antisymmetric tensor property n n

= w
r
w , where w
r
is the
axial vector associated with w. Expanding the above equation, we obtain:
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )
) 3 ( ) 3 ( ) 3 (
3
) 3 ( ) 3 ( ) 2 (
2
) 3 ( ) 3 ( ) 1 (
1
) 2 ( ) 2 ( ) 3 (
3
) 2 ( ) 2 ( ) 2 (
2
) 2 ( ) 2 ( ) 1 (
1
) 1 ( ) 1 ( ) 3 (
3
) 1 ( ) 1 ( ) 2 (
2
) 1 ( ) 1 ( ) 1 (
1
) 3 ( ) 3 ( ) ( ) 2 ( ) 2 ( ) ( ) 1 ( ) 1 ( ) (




n n n n n n n n n
n n n n n n n n n
n n n n n n n n n
n n n n n n n n n
+ + +
+ + + +
+ + + =
= + + =
w w w
w w w
w w w
w w w
b
b
b
b
b
b
w

On simplifying the above expression we obtain:
( ) ( )
( ) ( )
( ) ( )
) 3 ( ) 1 (
2
) 3 ( ) 2 (
1
) 2 ( ) 1 (
3
) 2 ( ) 3 (
1
) 1 ( ) 2 (
3
) 1 ( ) 3 (
2



n n n n
n n n n
n n n n
+
+ +
+ + =
w w
w w
w w w

Taking into account that
32 23 1
w w = = w ,
31 13 2
w w = = w ,
21 12 3
w w = = w , the
above equation becomes:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
76
) 3 ( ) 1 (
13
) 3 ( ) 2 (
23
) 2 ( ) 1 (
12
) 2 ( ) 3 (
32
) 1 ( ) 2 (
21
) 1 ( ) 3 (
31



n n n n
n n n n
n n n n
+ +
+ + +
+ + =
w w
w w
w w w

which is the same as:

=
=
3
1 ,
) ( ) (

b a
b a
b a
ab
n n w w
The terms V w and w V can be expressed as follows:

=
=

=
= =
|
|
.
|

\
|

|
|
|
.
|

\
|
=


3
1 ,
) ( ) (
3
1 ,
) ( ) ( ) ( ) (
3
1
) ( ) (
3
1 ,
) ( ) (


b a
b a
b a
ab b
b a
b a
b b b a
ab b
b
b b
b
b a
b a
b a
ab
n n n n n n
n n n n V
w w
w w

and

= =
=
|
|
|
.
|

\
|

|
|
.
|

\
|
=
3
1 ,
) ( ) (
3
1 ,
) ( ) (
3
1
) ( ) (

b a
b a
b a
ab a
b a
b a
b a
ab
a
a a
a
n n n n n n V w w w
Then,

=
=
|
|
|
.
|

\
|

|
|
|
.
|

\
|
=
3
1 ,
) ( ) (
3
1 ,
) ( ) (
3
1 ,
) ( ) (

) (

b a
b a
b a
a b ab
b a
b a
b a
ab a
b a
b a
b a
ab b
n n
n n n n V V
w
w w w w

Similarly, it is possible to show that:

=
=
3
1 ,
) ( ) ( 2 2 2 2

) (
b a
b a
b a
a b ab
n n V V w w w

1.5.6 Cayley-Hamilton Theorem
The Cayley-Hamilton theorem states that any tensor, T , satisfies its own characteristic
equation, i.e. if the eigenvalues of T satisfy the equation 0
2 3
= +
T T T
I I I I I I , so
does the tensor T :
0 1 T T T
T T T
= + I I I I I I
2 3
(1.309)
One of the applications of the Cayley-Hamilton theorem is to express the power of tensor,
n
T , as a combination of
1 n
T ,
2 n
T ,
3 n
T . For example,
4
T is obtained as:
T T T T 0 T 1 T T T T T T
T T T T T T
I I I I I I I I I I I I + = = +
2 3 4 2 3
(1.310)
Using the Cayley-Hamilton theorem, it is possible to express the third invariant as a
function of traces. According to the Cayley-Hamilton theorem, the expression
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
77
0 1 T T T
T T T
= + I I I I I I
2 3
remains valid. Additionally, by applying the double scalar
product with the second-order unit tensor, 1, we obtain:
1 0 1 1 1 T 1 T 1 T
T T T
: : : : : = + I I I I I I
2 3
(1.311)
Taking into consideration ) (
3 3
T 1 T Tr = : , ) (
2 2
T 1 T Tr = : , ) (T 1 T Tr = : , 3 ) ( = = 1 1 1 Tr : ,
0 ) ( = = 0 1 0 Tr : in the equation (1.311) we obtain:
[ ] ) ( ) ( ) (
3
1
0 ) ( ) ( ) ( ) (
2 3
3
2 3
T T T
1 T T T
T T T
T T T
Tr Tr Tr
Tr Tr Tr Tr
I I I I I I
I I I I I I
+ =
= +
=
3 2 1
(1.312)
Replacing the values of the invariants,
T
I ,
T
I I , given by equation (1.269), we obtain:
[ ]
)
`

+ =
3
2 3
) (
2
1
) ( ) (
2
3
) (
3
1
T T T T
T
Tr Tr Tr Tr I I I
(1.313)
or in indicial notation
)
`

+ =
kk jj ii kk ji ij ki jk ij
I I I T T T T T T T T T
2
1
2
3
3
1
T
(1.314)

Problem 1.35: Based on the Cayley-Hamilton theorem, find the inverse of a tensor T in
terms of tensor power.
Solution: The Cayley-Hamilton theorem states that:
0 1 T T T
T T T
= + I I I I I I
2 3

Carrying out the dot product between the previous equation and the tensor
1
T , we
obtain:
( ) 1 T T T
0 T 1 T T
T 0 T 1 T T T T T T
T T
T
T T T
T T T
I I I
I I I
I I I I I I
I I I I I I
+ =
= +
= +



2 1
1 2
1 1 1 1 2 1 3
1

The Cayley-Hamilton theorem also applies to square matrices of order n . Let
n n
A be a
square n by n matrix. The characteristic determinant is given by:
0 =

A
n n
1 (1.315)
where
n n
1 is the identity n by n matrix. Developing the determinant (1.315) we ontain:
0 ) 1 (
2
2
1
1
= +

n
n n n n
I I I L (1.316)
where
n
I I I , , ,
2 1
L are the invariants of A. In the particular case when 3 = n , the
invariants are the same obtained for a second-order tensor, i.e.:
A
I I =
1
,
A
I I I =
2
,
A
I I I I =
3
.
Applying the Cayley-Hamilton theorem it is true that:
0 1= + +

n
n n n n
I I I ) 1 (
2
2
1
1
L A A A (1.317)
By means of the relationship (1.317), we can obtain the inverse of the matrix
n n
A by
multiplying all the terms by the inverse,
1
A , i.e.:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
78
0 1
0 1
= + +
= + +



1
1
1 3
2
2
1
1
1 1 2
2
1 1
1
1
) 1 ( ) 1 (
) 1 (
A A A A
A A A A A A A
n
n
n
n n n n
n
n n n n
I I I I
I I I
L
L
(1.318)
then
( ) 1
1
1 3
2
2
1
1
1
1
) 1 (
) 1 (

=
n
n n n n
n
n
I I I
I
L A A A A (1.319)
n
I is the determinant of
n n
A . Then, the inverse exists if 0 ) ( = A det
n
I .
Problem 1.36: Check the Cayley-Hamilton theorem by using a second-order tensor whose
Cartesian components are given by:
(
(
(

=
1 0 0
0 2 0
0 0 5
T
Solution:
The Cayley-Hamilton theorem states that:
0 1 = +
T T T
I I I I I I T T T
2 3

where 8 1 2 5 = + + =
T
I , 17 5 2 10 = + + =
T
I I , 10 =
T
I I I , and
(
(
(

=
(
(
(

=
1 0 0
0 8 0
0 0 125
1 0 0
0 2 0
0 0 5
3
3
3
T ;
(
(
(

=
(
(
(

=
1 0 0
0 4 0
0 0 25
1 0 0
0 2 0
0 0 5
2
2
2
T
By applying the Cayley-Hamilton theorem, we can verify that it is true:
(
(
(

=
(
(
(

(
(
(

+
(
(
(

(
(
(

0 0 0
0 0 0
0 0 0
1 0 0
0 1 0
0 0 1
10
1 0 0
0 2 0
0 0 5
17
1 0 0
0 4 0
0 0 25
8
1 0 0
0 8 0
0 0 125

1.5.7 Norms of Tensors
The magnitude (module) of a tensor, also known as the Frobenius norm, is given below:
i i
v v = = v v v
r r r
(vector) (1.320)
ij ij
T T = = T T T : (second-order tensor)
(1.321)
ijk ijk
A A = = A A A : (third-order tensor)
(1.322)
ijkl ijkl
C C = = C C C :: (fourth-order tensor)
(1.323)
Interpreting the Frobenius norm of T is done by considering the principal space of T
where
3 2 1
, , T T T are the eigenvalues of T . In this space, it follows that:
T T
T T T I I I
ij ij
2
2 2
3
2
2
2
1
= + + = = = T T T T T : (1.324)
As we can verify T is an invariant, and T represents a measurement of distance as
shown in Figure 1.28.
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
79








Figure 1.28: Norm of a second-order tensor.
1.5.8 Isotropic and Anisotropic Tensor
A tensor is called isotropic when its components are the same in any coordinate system,
otherwise the tensor is said to be anisotropic.
Let T and T represent the tensor components T in the systems
i
e

and
i
e

,
respectively, so, the tensor is isotropic if T T = on any arbitrary basis.
Isotropic first-order tensor
Let v
r
be a vector that is represented by its components,
3 2 1
, , v v v , in the coordinate
system
3 2 1
, , x x x . The representation of these components in a new coordinate system,
3 2 1
, , x x x , are given by
3 2 1
, , v v v , so the transformation law for these components is:
j ij i j j i i
a v v v v = = = e e v

r
(1.325)
By definition, v
r
is an isotropic tensor if it holds that
i i
v v = , and this is only possible if
j i
e e =

, i.e. there is no change of system, or if the tensor is the zero vector, i.e.
i i i
0 v v = = . Then, the unique isotropic first-order tensor is the zero vector 0
r
.
Isotropic second-order tensor
An example of a second-order isotropic tensor is the unit tensor, 1, whose components
are represented by
kl
(Kronecker delta). In the demonstration, we use the transformation
law for a second-order tensor components, obtained in (1.248), thus:
ij
T
jk ik kl jl ik ij
a a a a = = =
=

3 2 1
1 AA

(1.326)
An immediate observation of the isotropy of unit tensor 1 is that any spherical tensor
( 1 ) is also an isotropic tensor. So, if a second-order tensor is isotropic it is spherical and
vice versa.
Isotropic third-order tensor
An example of a third-order isotropic tensor is the Levi-Civita pseudo-tensor, defined in
(1.182), which is not a real tensor in the strict meaning of the word. With reference to
the transformation law for the third-order tensor components, (see equation (1.248)), we
can conclude that:

2
x

1
x

3
x

1
T

2
T

3
T
T

T T
T T T I I I 2
2
= = :
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
80
{

1
ijk ijk lmn kn jm il ijk
a a a = = = A (see Problem 1.21)
(1.327)
Isotropic fourth-order tensor
With reference to the transformation law for fourth-order tensor components, (see
equation (1.249)), it is possible to demonstrate that the following tensors are isotropic:
jk il ijkl jl ik ijkl kl ij ijkl
= = = I I I ; ; (1.328)
Therefore, any fourth-order isotropic tensor can be represented by a linear combination of
the three tensors given in (1.328), e.g.:
I I I D
2 1 0
a a a + + =
1 1 1 1 1 1 + + =
2 1 0
a a a D
jk il jl ik kl ij ijkl
a a a
2 1 0
+ + = D
(1.329)
Problem 1.37: Let C be a fourth-order tensor, whose components are given by:
( )
jk il jl ik kl ij ijkl
+ + = C
where , are constant real numbers. Show that C is an isotropic tensor.
Solution:
Applying the transformation law for fourth-order tensor components:
mnpq lq kp jn im ijkl
a a a a C C =
and by replacing the relation ( )
np mq nq mp pq mn mnpq
+ + = C in the above equation,
we obtain:
( ) [ ]
( )
( )
( )
ijkl
jk il jl ik kl ij
lq kn jn iq lq kp jq ip lq kq jn in
np mq lq kp jn im nq mp lq kp jn im pq mn lq kp jn im
np mq nq mp pq mn lq kp jn im ijkl
a a a a a a a a a a a a
a a a a a a a a a a a a
a a a a
C
C
=
+ + =
+ + =
+ + =
+ + =





which is proof that C is an isotropic tensor.

1.5.9 Coaxial Tensors
Two arbitrary second-order tensors, T and S , are coaxial tensors if they have the same
eigenvectors. It is easy to show that if two tensors are coaxial, this means the dot product
between them is commutative, and vice versa, i.e.:
coaxial are if , T S T S S T =
(1.330)
If T and S are coaxial as well as symmetric tensors, the spectral representations of these
tensors are given by:

= =
= =
3
1
) ( ) (
3
1
) ( ) (

;

a
a a
a
a
a a
a
n n S n n T S T (1.331)
An immediate result of (1.330) is that the tensor S and its inverse
1
S are coaxial tensors:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
81

=

=

= =
= =
3
1
) ( ) ( 1
3
1
) ( ) (
1 1

1
;

a
a a
a a
a a
a
n n S n n S
1 S S S S
S
S
(1.332)
where
a
S ,
a
S
1
, are the eigenvalues of S and
1
S , respectively.
If S and T are coaxial symmetric tensors, the resulting tensor ( T S ) becomes another
symmetric tensor. To prove this we start from the definition of coaxial tensors:
0 S T 0 S T S T 0 T S S T T S S T = = = =
skew T
) ( 2 ) ( (1.333)
Then, if the antisymmetric part of a tensor is a zero tensor, it follows that this tensor is
symmetric:
sym skew
) ( ) ( ) ( S T S T 0 S T = (1.334)
1.5.10 Polar Decomposition
Let F be an arbitrary nonsingular second-order tensor, i.e. (
1
0 ) (

F F det ).
Additionally, as previously seen, it satisfies the condition 0 n n N
n
N N
r r r
= = =

(
)

( )

(
f f F ,
since 0 ) ( F det . After that, given an orthonormal basis
) (

a
N , we can obtain:

=
= =
=

=
= = =
= =

3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) ( 1



a
a a
a
a
a a
a
a a
a
a a
N n
N N N N 1
N N 1
F
F F F F
F F
(1.335)
NOTE: The representation of F , given in (1.335), is not the spectral representation of F
in the strict sense of the word, i.e.,
a
are not eigenvalues of F , and neither
) (

a
n nor
) (

a
N
are eigenvectors of F .










Figure 1.29: Projecting F onto
) (

a
N .

) 2 (

n

) 1 (

n

) 3 (

N

) 1 (

N

) 2 (

N

) 3 (

n

) 1 ( )

(

) 1 (
N
N
= F f
r


) 2 ( )

(

) 2 (
N
N
= F f
r


) 3 ( )

(

) 3 (
N
N
= F f
r


) 3 (
3
) 3 ( )

( )

( ) 3 (
) 2 (
2
) 2 ( )

( )

( ) 2 (
) 1 (
1
) 1 ( )

( )

( ) 1 (

) 3 ( ) 3 (
) 2 ( ) 2 (
) 1 ( ) 1 (
n n N
n n N
n n N
N N
N N
N N
= = =
= = =
= = =

f f F
f f F
f f F
r r
r r
r r

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
82

Note that for the arbitrary orthonormal basis
) (

a
N , the new basis
) (

a
n will not necessarily
be orthonormal. We seek to find a basis
) (

a
N so that the new basis
) (

a
n is orthonormal,
(see Figure 1.29), i.e. 0
)

( )

(
) 2 ( ) 1 (
=
N N
f f
r r
, 0
)

( )

(
) 3 ( ) 2 (
=
N N
f f
r r
, 0
)

( )

(
) 1 ( ) 3 (
=
N N
f f
r r
. Then we
look for a space in accordance with the following orthogonal transformation
) ( ) (

a a
N R n = , which ensures
) (

a
n orthonormality since an orthogonal transformation
changes neither angles between vectors nor their magnitudes.
Now, consider that there is a transformation from
) (

a
N to
) (

a
n , which is given by the
following orthogonal transformation
) ( ) (

a a
N R n = , then we can state that:
F F
F


= =
= = = =

= = =
T
a
a a
a
a
a a
a
a
a a
a
R U U R
U R N N R N N R N n
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (

(1.336)
where we have defined the tensor

=
=
3
1
) ( ) (

a
a a
a
N N U . Note that U is a symmetric
tensor, i.e.
T
U U= . This condition is easily verified by the fact that
) ( ) (

a a
N N is also
symmetric. Now considering that R n n R N N R n = = =
) ( ) ( ) ( ) ( ) (

a a T a a a
, we obtain:
T
a
a a
a
a
a a
a
a
a a
a
R V R V
R V R n n R n n N n


= =
= = = =

= = =
F F
F
3
1
) ( ) (
3
1
) ( ) (
3
1
) ( ) (

(1.337)
where we have defined the symmetric second-order tensor

=
=
3
1
) ( ) (

a
a a
a
n n V . By
comparing the spectral representation of U with V , we can conclude that they have the
same eigenvalues but different eigenvectors, and they are related by
) ( ) (

a a
N R n = .
With reference to the above considerations, we can define the polar decomposition:
R V U R = = F Polar Decomposition (1.338)
Carrying out the dot product between
T
F and U R = F , we obtain:
C F F F F F F
T T T T T
C
T
= = = = = = U U U U U R U R ) (
2
3 2 1

(1.339)
Moreover, by carrying out the dot product between R V = F and
T
F , we obtain:
b F F F F F F
T T T T T
b
T
= = = = = = V V V V R V R V ) (
2
3 2 1

(1.340)
Since 0 ) ( F det , the tensors C and b are positive definite symmetric tensors, (see
Problem 1.25), which implies that the eigenvalues of C and b are all real and positive.
However, up to now, 0 ) ( F det is the only restriction imposed on the tensor F .
Therefore, we have the following possibilities:
If 0 ) ( > F det
In this scenario, we have 0 ) ( ) ( ) ( ) ( ) ( > = = R V U R det det det det det F , which results in the
following cases:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
83

tensors definite Positive ,


tensor ortogonal Proper
V U
R
or

tensors definite Negative ,


tensor orthogonal Improper
V U
R

If 0 ) ( < F det
In this situation, we have 0 ) ( ) ( ) ( ) ( ) ( < = = R V U R det det det det det F , which give us the
following cases:

tensors definite Negative ,


tensor orthogonal Proper
V U
R
or

tensors definite Positive ,


tensor orthogonal Improper
V U
R

NOTE: In Chapter 2 we will work with some special tensors where F is a nonsingular
tensor, 0 ) ( F det , and 0 ) ( > F det . U and V are positive definite tensors and R is a
rotation tensor, i.e. a proper orthogonal tensor.
1.5.11 Partial Derivative with Tensors
The first derivative of a tensor with respect to itself is defined as:
I = =

)

( )

( ,
l k j i jl ik l k j i
kl
ij
e e e e e e e e A
A
A
A

A
A
(1.341)
The derivative of a tensor trace with respect to a tensor:
[ ]
[ ] 1 e e e e e e A
A
A
A
= = =

)

( )

( )

( ) (
) (
, j i ij j i kj ki j i
ij
kk

A
A
Tr
Tr

(1.342)
The derivative of the tensor trace squared with respect to the tensor is given by:
[ ] [ ]
1 A
A
A
A
A
A
) ( 2
) (
) ( 2
) (
2
Tr
Tr
Tr
Tr
=

(1.343)
And, the derivative of the trace of the tensor squared with respect to tensor is given by:
[ ]
[ ] [ ]
T
j i ji
j i ji ji j i sj ri sr rj si rs
j i
ij
rs
sr
ij
sr
rs j i
ij
rs sr
A e e
e e e e
e e e e
A
A
2 )

( 2
)

( )

(
)

(
) ( ) (
)

(
) ( ) (
2
= =
+ = + =

(
(

A
A A A A
A
A
A
A
A
A
A
A A Tr

(1.344)
We leave the reader with the following demonstration:
[ ]
T
) ( 3
) (
2
3
A
A
A
=

Tr

(1.345)
Then, if we are considering a symmetric second-order tensor, C, it is true that
[ ]
1
C
C
=

) ( Tr
,
[ ]
1 C
C
C
) ( 2
) (
2
Tr
Tr
=

,
[ ]
C C
C
C
2 2
) (
2
= =

T
Tr
,
[ ]
2 2
3
3 ) ( 3
) (
C C
C
C
= =

T
Tr
.
Moreover, we can say that the derivative of the Frobenius norm of C is given by:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
84
( )
[ ] [ ]
[ ] C C
C C
C
C
C
C C
C
C C
C
C
C
2 ) (
2
1
) ( ) (
2
1
) ( ) (
2
1
2
2
2
1
2
2
,

=
=

|
.
|

\
|

|
.
|

\
|


Tr
Tr Tr
Tr Tr
T
:
(1.346)
or:
C
C
C
C
C
C
= =

) (
2
Tr
(1.347)
Another interesting derivative is presented below:
( )
j kj j
sym
kj j jk kj j jk j kj
ik i j kj jk ij i j ij ik
k
j
ij i j ij
k
i
k
j ij i
n C n C n C C n C n C
C n n C C n n C
n
n
C n n C
n
n
n
n C n
2 2 ) ( = = + = + =
+ = + =



(1.348)
where we have assumed that C is symmetric, i.e.,
jk kj
C C = .
Let C be a symmetric second-order tensor. The partial derivative of
1
C with respect to
the tensor C is obtained by using the following relationship:
( )
O =

C
C C
C
1
1
(1.349)
where O is the fourth-order zero tensor and the above equation in indicial notation
becomes:
( ) ( ) ( )
( ) ( ) ( ) ( )
( ) ( )
1 1
1
1 1 1
1
1
1
1
1 1

jr
kl
qj
iq qr
kl
iq
jr
kl
qj
iq jr qj
kl
iq
kl
qj
iq qj
kl
iq
ikjl
kl
qj
iq qj
kl
iq
kl
qj iq
C
C
C
C
C
C
C
C
C
C C C
C
C
C
C
C C
C
C
C
C
C C
C
C
C
C C

O
(1.350)
whereas ( )
jq qj qj
C C C + =
2
1
, so we can conclude that:
( ) ( )
( )
[ ] [ ]
( )
[ ]
1 1 1 1
1
1 1 1 1 1 1
1
1 1
1
2
1
2
1
2
1
2
1

+ =

+ = + =

+
=

kr il lr ik
kl
ir
jr ql jk iq jr jl qk iq jr ql jk jl qk iq
kl
ir
jr
kl
jq qj
iq qr
kl
iq
C C C C
C
C
C C C C C C
C
C
C
C
C C
C
C
C

(1.351)
Or in tensorial notation:
[ ]
1 1 1 1
1
2
1

+ =

C C C C
C
C
(1.352)
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
85
NOTE: Note that, if we had not replaced the symmetric part of
qj
C in (1.351), we would
have found that
( ) ( )
1 1 1 1 1 1
1

= =

lr ik jr jl qk iq jr
kl
qj
iq qr
kl
iq
C C C C C
C
C
C
C
C
, which is a non-
symmetric tensor.
1.5.11.1 Partial Derivative of Invariants
Let T be a second-order tensor. The partial derivative of
T
I with respect to T , (see
equation (1.342)), is:
[ ] [ ]
[ ] 1 T
T
T
T
T
T
= =

,
) (
) (
Tr
Tr I
(1.353)
The partial derivative of
T
I I with respect to T , (see equation (1.342)), is:
[ ]
[ ] [ ]
[ ] [ ]
[ ]
T
T
I I
T 1 T
T 1 T
T
T
T
T
T T
T T
T
=
=
(
(

=
)
`

) (
2 ) ( 2
2
1
) ( ) (
2
1
) ( ) (
2
1
2 2
2 2
Tr
Tr
Tr Tr
Tr Tr

(1.354)
Next, we apply the Cayley-Hamilton theorem so as to represent T as:
2 1
2 1
2 2 2 2 2 3



+ =
= +
= +
T T 1 T
0 T T 1 T
0 T 1 T T T T T T
T T T
T T T
T T T
I I I I I I
I I I I I I
I I I I I I : : : :
(1.355)
By substituting (1.355) into the equation in (1.354), we obtain:
[ ]
( ) ( )
T T
T
I I I I I I I I I I I
I I
2 1 2 1
) ( ) (

= + = =

T T T T 1 1 T T 1 T
T
T T T T T
T
Tr Tr (1.356)
To find the partial derivative of the third invariant, we can start with the definition given in
(1.313), so:
[ ]
[ ]
[ ] [ ]
[ ]
[ ]
[ ] [ ]
1 T T
1 T T T T T
1 T 1 T T T T
1 T
T
T
T T
T
T
T
T T T T
T T
T T
T
I I I
I I I
T T
T T
T T
T
+ =
+ =
+ =
+

=
)
`

) (
) ( ) (
2
1
) ( ) (
) (
2
1
) (
2
1
) ( ) (
) (
6
3 ) (
) (
2
1
) (
) (
2
1
) ( 3
3
1
) (
6
1
) ( ) (
2
1
) (
3
1
2
2
2
2
2
2 2
2
2
2
2
3
2 3
Tr Tr Tr
Tr Tr Tr
Tr
Tr
Tr Tr
Tr
Tr Tr Tr Tr
(1.357)
Once again using the Cayley-Hamilton theorem we obtain:
1 T T T
0 T 1 T T
0 T 1 T T T T T T
T T T
T T T
T T T
I I I I I I
I I I I I I
I I I I I I
+ =
= +
= +



2 1
1 2
1 1 1 2 1 3
(1.358)
and the transpose:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
86
( ) ( ) ( ) 1 T T 1 T T T
T T T T T
I I I I I I I I I
T
T T T
+ = + =
2 2 1

(1.359)
By comparing (1.357) with (1.359) we find another way to express the derivative of
T
I I I
with respect to T , i.e.:
[ ]
( )
T
T
I I I I I I
I I I

= =

T T
T
T T
T 1
(1.360)
1.5.11.2 Time Derivative of Tensors
Let us assume a second-order tensor depends on the time, t , i.e. ) t T( T = . Then, we define
the first time derivative and the second time derivative of the tensor T , respectively, as:
T T T T
& & &
= =
2
2
;
Dt
D
Dt
D
(1.361)
The time derivative of a tensor determinant is defined as:
( ) [ ] ( )
ij
ij
Dt
D
Dt
D
T cof
T
det = T (1.362)
where ( )
ij
T cof is the cofactor of
ij
T and defined as ( ) [ ] ( ) ( )
ij
T
ij
1

= T T det T cof .

Problem 1.38: Consider that [ ] ( )2
1
2
1
) (
b
b I I I J = = det , where b is a symmetric second-order
tensor, i.e.
T
b b = . Obtain the partial derivatives of J and ) (J ln with respect to b .

Solution:
( )
( ) ( )
( )
1 1
2
1
2
1
2
1
2
1

2
1

2
1
2
1
2
1


= =
=

=

(

b b
b
b
b b
b
b b
b
b
b
J I I I
I I I I I I
I I I
I I I
I I I
J
T

( ) [ ]
1
2
1

2
1
2
1

=

|
.
|

\
|

b
b b b
b
b
b
I I I
I I I
I I I
J
ln
ln

1.5.12 Spherical and Deviatoric Tensors
Any tensor can be decomposed into a spherical and a deviatoric part, so, for a given
second-order tensor T , this decomposition is represented by:
dev
m
dev dev dev sph
I
T 1 T 1 T 1
T
T T T
T
+ = + = + = + = T
Tr
3 3
) (
(1.363)
The deviatoric part of the tensor T is defined as:


1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
87
1 T 1
T
T T
m
dev
T
Tr
= =
3
) (
(1.364)
For the following operations, we consider that T is a symmetric tensor,
T
T T = , then
under this condition the deviatoric tensor components,
dev
ij
T , become:
(
(
(




=
(
(
(

=
(
(
(

=
) 2 (
) 2 (
) 2 (
22 11 33 3
1
23 13
23 33 11 22 3
1
12
13 12 33 22 11 3
1
33 23 13
23 22 12
13 12 11
33 23 13
23 22 12
13 12 11
T T T T T
T T T T T
T T T T T
T T T T
T T T T
T T T T
T T T
T T T
T T T
T
m
m
m
dev dev dev
dev dev dev
dev dev dev
dev
ij

(1.365)
Graphical representations of the Cartesian components of the spherical and deviatoric
parts are shown in Figure 1.30.
In the following subsections we obtain the deviatoric tensor invariants in terms of the
principal invariants of T .
1.5.12.1 First Invariant of the Deviatoric Tensor
0 ) (
3
) (
) (
3
) (
) (
3
= =
(

= =
=
3 2 1
ii
dev
dev
I

1
T
T 1
T
T T
T
Tr
Tr
Tr
Tr
Tr Tr
(1.366)
Thus, we can conclude that the trace of any deviatoric tensor is equal to zero.
1.5.12.2 Second Invariant of the Deviatoric Tensor
For simplicity we can use the principal space to obtain the second and third invariant of the
deviatoric tensor. In the principal space the components of T are given by:
(
(
(

=
3
2
1
0 0
0 0
0 0
T
T
T
T
ij

(1.367)
The principal invariants of T :
3 2 1
T T T + + =
T
I ,
1 3 3 2 2 1
T T T T T T + + =
T
I I ,
3 2 1
T T T =
T
I I I .
The deviatoric components, 1 T T
m
dev
T = , in the principal space are:
(
(
(

=
m
m
m
dev
ij
T T
T T
T T
T
3
2
1
0 0
0 0
0 0

(1.368)
So, the second invariant of deviatoric tensor
dev
T is evaluated as follows:
( )
2
2
2
3 2 1 3 2 3 1 2 1
3 2 3 1 2 1
3
3
1
3
) (
3
2
3 ) ( 2 ) (
) )( ( ) )( ( ) )( (
T T
T
T
T
T
T
I I I
I
I
I
I I
I I
m m
m m m m m m dev
=
+ =
+ + + + + =
+ + =
T T T T T T T T T T T
T T T T T T T T T T T T
(1.369)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
88
We could also have obtained the above result, by directly starting from the definition of the
second invariant of a tensor given in (1.269), i.e.:
[ ] { } [ ] { }
[ ] { }
[ ] { }
[ ]
(

+ =
(

+ =
+ =
+ =
=
= =

3
) (
2
1
3
9 3
2 ) (
2
1
) ( ) ( 2 ) (
2
1
) 2 (
2
1
) (
2
1
) (
2
1
) ( ) (
2
1
2
2
2
2
2 2
2 2
2
2 2 2
T
T
T
T
T
T
1 T T
1 1 T T
1 T
T T
T T
I
I
I
I
I I I
m m
m m
m
dev dev
dev dev
Tr
Tr
Tr T Tr T Tr
T T Tr
T Tr
Tr Tr
(1.370)
Observing that
T
T
T I I I 2 ) (
2 2
3
2
2
2
1
2
= + + = T T T Tr , (see Problem 1.31), the equation
(1.370) becomes:
( )
2
2 2
2
3
3
1
3
2
2
2
1
3
2
2
1
T T
T T
T
T T
T
I I I
I
I I
I
I I I I I
dev
=
(

=
(

+ + = (1.371)




















Figure 1.30: Spherical and deviatoric part.

4 4 4 4 4 4 4 4 3 4 4 4 4 4 4 4 4 2 1


11
T

12
T
13
T

33
T

23
T

13
T

22
T

23
T

12
T

1
x

2
x

m
T

m
T

m
T

1
x

2
x

3
x

dev
11
T

12
T
13
T

dev
33
T

23
T

13
T

dev
22
T

23
T

12
T

1
x

2
x

3
x
+

3
x
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
89
Another equation for
dev
I I
T
is presented in terms of deviatoric tensor components. To
calculate this, we can apply the equation (1.370):
[ ] [ ]
dev
ji
dev
ij
dev dev dev dev dev
dev
I I T T Tr Tr
2
1
2
1
) (
2
1
) (
2
1
2
= = = = T T T T T
T
(1.372)
Expanding the previous equation we obtain:
[ ]
2
23
2
13
2
12
2
33
2
22
2
11
) ( 2 ) ( 2 ) ( 2 ) ( ) ( ) (
2
1
dev dev dev dev dev dev
dev
I I T T T T T T + + + + + =
T
(1.373)
Additionally, in the space of the principal directions we obtain:
[ ]
2
3
2
2
2
1
) ( ) ( ) (
2
1
2
1
dev dev dev dev
ji
dev
ij dev
I I T T T T T + + = =
T
(1.374)
Another way to express the second invariant is shown below:
[ ]
2
13
2
23
2
12 22 11 33 11 33 22
22 12
12 11
33 13
13 11
33 23
23 22
) ( ) ( ) ( 2 2 2
2
1
dev dev dev dev dev dev dev dev dev
dev dev
dev dev
dev dev
dev dev
dev dev
dev dev
dev
I I
T T T T T T T T T
T T
T T
T T
T T
T T
T T
=
+ + =
T
(1.375)
or
( ) ( ) ( ) ( )
( ) ( ) ( ) ( ) ( )
2
13
2
23
2
12
2
33
2
22
2
11
2
22 22 11
2
11
2
33 33 11
2
11
2
33 33 22
2
22
) ( ) ( ) (
2
2 2
2
1
dev dev dev
dev dev dev dev dev dev dev
dev dev dev dev dev dev dev dev
dev
I I
T T T
T T T T T T T
T T T T T T T T

+ + +
(

(
+

+ + + + =
T

(1.376)
Note that, from equation (1.373), we can state that:
2
23
2
13
2
12
2
33
2
22
2
11
) ( 2 ) ( 2 ) ( 2 2 ) ( ) ( ) (
dev dev dev
dev
dev dev dev
I I T T T T T T = + +
T

(1.377)
Substituting (1.377) into (1.376), we find:
[ ]
2
13
2
23
2
12
2
22 11
2
33 11
2
33 22
) ( ) ( ) ( ) ( ) ( ) (
6
1
dev dev dev dev dev dev dev dev dev
dev
I I T T T T T T T T T + + =
T


(1.378)
Moreover, if we consider the principal space we obtain:
[ ]
2
2 1
2
3 1
2
3 2
) ( ) ( ) (
6
1
dev dev dev dev dev dev
dev
I I T T T T T T + + =
T
(1.379)
1.5.12.3 Third Invariant of Deviatoric Tensor
The third invariant of the deviatoric tensor is given by:
( )
T T T T
T T T
T
T
T
T
T
T
T
T
I I I I I I I
I I I I
I I I
I
I
I
I I
I
I I I
I I I
m m m
m m m dev
27 9 2
27
1
27
2
3
27 9 3
) ( ) (
) )( )( (
3
3
3 2
3
3 2 1
2
3 2 3 1 2 1 3 2 1
3 2 1
+ =
+ =
+ =
+ + + + + =
=
T T T T T T T T T T T T T T T
T T T T T T
(1.380)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
90
Another way of expressing the third invariant is:
dev
ki
dev
jk
dev
ij
dev dev dev
dev
I I I T T T T T T
3
1
3 2 1
= =
T
(1.381)
Problem 1.39: Let be a symmetric second-order tensor, and
dev
s be a deviatoric
tensor. Prove that s

s
s =

: . Also show that and


dev
are coaxial tensors.
Solution: First, we make use of the definition of a deviatoric tensor:
1 s s 1 s

3 3
I I
sph dev sph
= + = + = + = .
Afterwards we calculate:
[ ] [ ]
1

I
I
3
1 3

which in indicial notation is:
[ ]
ij kl jl ik ij
kl kl
ij
kl
ij
I

3
1
3
1
=


s

Therefore
{
kl
ii kl kl ij kl ij jl ik ij ij kl jl ik ij
kl
ij
ij
s
s s s s s
s
s
=
= = |
.
|

\
|
=

=0
3
1
3
1
3
1


s

s
s =

:
To show that two tensors are coaxial, we must prove that
dev dev
= :
1
1 1
1




= |
.
|

\
|
=
= =
= = =
dev
sph sph dev
I
I I
I
3
3 3
3
) (

Therefore, we have shown that and
dev
are coaxial tensors. In other words, they have
the same principal directions.

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
91
1.6 The Tensor-Valued Tensor Function
Tensor-valued tensor function can be of the types: scalar, vector, or higher-order tensors.
As examples of scalar-valued tensor functions we can list:
( )
S T S T
T T
: = =
= =
) , (
) (

det
(1.382)
where T and S are second-order tensors. Additionally, as an example of a second-order-
valued tensor function we have:
( ) T 1 T + = =
(1.383)
where and are scalars.
1.6.1 The Tensor Series
The function ) (x f can be approximated by the Taylor series as
n
n
n
n
a x
x
a f
n
x f ) (
) (
!
1
) (
0

=
, where ! n denotes the factorial of n , and ) (a f is the value of
the function at the application point a x = . We can extrapolate that definition for use on
tensors. For example, let us suppose we have a scalar-valued tensor function in terms of
a second-order tensor, E , then we can approximate ) (E as:
L
L
+

+
+

+
) (
) (
) (
2
1
) (
) (
) )( (
) (
! 2
1
) (
) (
! 1
1
) (
! 0
1
) (
0
0
2
0 0
0
0
0
0
0
2
0
0
0
E E
E E
E
E E E E
E
E
E E
E E
: : :



kl kl ij ij
kl ij
ij ij
ij
E E E E
E E
E E
E
(1.384)
A second-order-valued tensor function, ) (E S , can be approximated as:
L
L
+

+
+

+
) (
) (
) (
2
1
) (
) (
) (
) (
) (
! 2
1
) (
) (
! 1
1
) (
! 0
1
) (
0
0
2
0 0
0
0
0
0
2
0 0
0
0
E E
E E
E
E E E E
E
E
E E
E E
E
E E E E
E
E
E E
: : :
: : :
S S
S
S S
S S
(1.385)
Other tensor algebraic expressions can be represented by series, e.g.:
L
L
L
+ =
+ = +
+ + + + =
5 3
3 2
3 2
! 5
1
! 3
1
) sin(
3
1
2
1
) (
! 3
1
! 2
1
S S S S
S S S S 1
S S S 1
S
ln
exp
(1.386)
With reference to the spectral representation of a symmetric second-order tensor, S , it is
also true that:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
92


= =
= =
+ = |
.
|

\
|
+ + = +
=
|
|
.
|

\
|
+

+ + =

3
1
) ( ) (
3
1
) ( ) ( 3 2
3
1
) ( ) (
3
1
) ( ) (
3 2

) 1 (

3
1
2
1
) (

! 3 ! 2
1
a
a a
a
a
a a
a a a
a
a a
a
a
a a a a
a
n n n n S 1
n n n n
S
ln ln
exp exp
L
L

(1.387)
where
a
and
) (

a
n are the eigenvalues and eigenvectors, respectively, of the tensor S .
1.6.2 The Tensor-Valued Isotropic Tensor Function
A second-order-valued tensor function, ) (T = , is isotropic if after an orthogonal
transformation the following condition is satisfied:
( ) ( ) ( )
( )
4 43 4 42 1
*
*
T
Q T Q Q T Q T

= =
T T

(1.388)
We can show that ) (T has the same principal directions of T , i.e. ) (T and T are
coaxial tensors. To demonstrate this we can regard the components of T in the principal
space as:
( )
(
(
(

=
3
2
1
0 0
0 0
0 0
ij
T
(1.389)
Then the tensor function is given in terms of the principal values of T : ( )
3 2 1
, , =
and, the transformation of T is given by:
T
Q T Q T =
*
(1.390)
Likewise, for the tensor function :
( ) ( )
T
Q T Q T =
*
(1.391)
If we take as the orthogonal tensor components:
( )
(
(
(

=
1 0 0
0 1 0
0 0 1
ij
Q
(1.392)
After having done the calculation for the matrices (1.391), we obtain:
(
(
(

=
=
(
(
(




=
(
(
(




=
33
22
11
*
33 23 13
23 22 12
13 12 11
33 23 13
23 22 12
13 12 11
*
0 0
0 0
0 0
(1.393)
To satisfy that =
*
(isotropy), we conclude that 0
23 13 12
= = = . Therefore, ) (T
and T have the same principal directions.
Once again we observe, a tensor function ) (T . This tensor function is isotropic if and
only if it can be represented by the following linear transformation, Truesdell & Noll
(1965):
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
93
2
2 1 0
) ( T T 1 T + + = = (1.394)
where
0
,
1
,
2
are functions of the tensor T invariants or functions of the T
eigenvalues.
A brief demonstration follows. We can now consider the spectral representations of T and
, respectively:
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1
3
1
) ( ) (

n n n n n n n n T + + = =

= a
a a
a
(1.395)
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1
3
1
) ( ) (

n n n n n n n n + + = =

= a
a a
a
(1.396)
Note that T and have the same principal directions
) (

i
n . Then, we can put the
following set of equations together:

+ + =
+ + =
+ + =
) 3 ( ) 3 ( 2
3
) 2 ( ) 2 ( 2
3
) 1 ( ) 1 ( 2
1
2
) 3 ( ) 3 (
3
) 2 ( ) 2 (
2
) 1 ( ) 1 (
1
) 3 ( ) 3 ( ) 2 ( ) 2 ( ) 1 ( ) 1 (



n n n n n n T
n n n n n n T
n n n n n n 1
(1.397)
Solving the set above, we obtain
) ( ) ( ) (

a a a
M n n as a function of the tensor T , and we
obtain:
( )
( )
( )
) )( ( ) )( ( ) )( (
) )( ( ) )( ( ) )( (
) )( ( ) )( ( ) )( (
2 3 1 3
2
2 3 1 3
2 1
2 3 1 3
2 1 ) 3 (
3 2 1 2
2
3 2 1 2
3 1
3 2 1 2
3 1 ) 2 (
2 1 3 1
2
2 1 3 1
3 2
2 1 3 1
3 2 ) 1 (

+

+



=

+

+



=

+

+



=
T
T 1 M
T
T 1 M
T
T 1 M
(1.398)
It is evident that, if we substitute the values of
) ( ) ( ) (

a a a
M n n in equation (1.395) we
obtain: T T = . Now, if we substitute the values of
) ( ) ( ) (

a a a
M n n in equation (1.396),
we obtain:
2
2 1 0
) ( T T 1 T + + = = (1.399)
where the coefficients
0
,
1
, and
2
are functions of the eigenvalues of T ,
) (
3 2 1
, and given by:
( ) ( ) ( )
) )( ( ) )( ( ) )( (
) )( ( ) )( ( ) )( (
) )( ( ) )( ( ) )( (
2 3 1 3
3
3 2 1 2
2
2 1 3 1
1
2
2 3 1 3
2 1 3
3 2 1 2
3 1 2
2 1 3 1
3 2 1
1
2 3 1 3
2 1 3
3 2 1 2
3 1 2
2 1 3 1
3 2 1
0


+


+


=

+


+
=


+


+


=

(1.400)
We can now show that if a tensor function ) (T is given in (1.399), this tensor function is
isotropic:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
94
( ) ( ) ( )
) (
*
2
*
2
*
1 0
2
2 1 0
2
2 1 0
*
T
T T 1 Q T Q Q T Q Q 1 Q
Q T T 1 Q Q T Q T
=
+ + = + + =
+ + = =


T T T
T T

(1.401)
1.6.3 The Derivative of the Tensor-Valued Tensor Function
Firstly, we refer to a scalar-valued tensor function:
) (A =
(1.402)
The partial derivative of ) (A with respect to A is defined as:
)

( ,
j i
ij
e e
A
A


A

(1.403)
where the comma denotes a partial derivative.
Then the second derivative of ) (A becomes a fourth-order tensor:
)

( )

( ,
2 2
l k j i ijkl l k j i
kl ij
e e e e e e e e
A A
AA
=


= =


D
A A

(1.404)
Let C and b be positive definite symmetric second-order tensors defined as:
T T
F F b F F C = = ; (1.405)
where F is an arbitrary second-order tensor with the restriction 0 ) ( > F det imposed on it.
We must also bear in mind that there is a scalar-valued isotropic tensor function,
( )
C C C
I I I I I I , , = , expressed in terms of the principal invariants of C , where
b C
I I = ,
b C
I I I I = ,
b C
I I I I I I = . Next, we can find the partial derivative of with respect to C , and
with respect to b . We must also verify that the following relation holds:
b F F
b C
= , ,
T
(1.406)
By applying the chain rule for derivative we obtain:
( )
C C C C
C
C
C
C
C
C
C C C
C

=
I I I
I I I
I I
I I
I
I
I I I I I I

, ,
,
(1.407)
Considering the partial derivatives of the principal invariants, we can state that:
1 =

C
C
I

2 1
= = =

C C C C
C
C C C C
C
I I I I I I I
I I
T
1 1
1
C C C C
C
C C C C
C
I I I I I I I I I
I I I
T
+ = = =

2 1

(1.408)
Now, by substituting the following values 1 =

C
C
I
, C
C
C
C
=

1 I
I I
and
1
=

C
C
C
C
I I I
I I I

into the equation in (1.407), we obtain:
( )
1
,

= C C
C
C
C
C C
C
I I I
I I I
I
I I I

1 1
(1.409)
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
95
1
,

|
|
.
|

\
|

+
|
|
.
|

\
|

|
|
.
|

\
|

= C C
C
C C
C
C C
C
I I I
I I I I I
I
I I I

1 (1.410)
Another way to express the relation (1.410) is by substituting 1 =

C
C
I
, C
C
C
C
=

1 I
I I

and 1
C C
C
C C
C
I I I
I I I
+ =

2
into the equation in (1.407), thus:
2
, C C
C
C
C C
C
C
C
C C
C
|
|
.
|

\
|

+
|
|
.
|

\
|

|
|
.
|

\
|

=
I I I
I
I I I I I
I I
I I I
I
I I I

1 (1.411)
If we now consider the values 1 =

C
C
I
,
2 1
=

C C
C
C C
C
I I I I I
I I
,
1
=

C
C
C
C
I I I
I I I
, in the
equation in (1.407), we obtain:
2 1
,

|
|
.
|

\
|

|
|
.
|

\
|

+
|
|
.
|

\
|

= C C
C
C
C
C
C
C C
C
I I I
I I
I I I
I I I
I I
I I I

1 (1.412)
If we now observe both
b C
I I = ,
b C
I I I I = ,
b C
I I I I I I = , and the equation in (1.410), we can
draw the conclusion that:
1
,

|
|
.
|

\
|

= b b
b
b b
b
b b
b
I I I
I I I I I
I
I I I

1 (1.413)
Using the equation in (1.410), the equation
T
F F
C

,
becomes:
T T T T
I I I
I I I I I
I
I I I
F C F F C F F F F F
C
C C
C
C C
C

|
|
.
|

\
|

=
1
,

1
(1.414)
Then, if we observe that:
b F F F F = =
T T
1
2
b b b F F F F F C F
F F C
= = =
=


T T T
T

(1.415)
b b F F b F F F C F
F b F C




= =
=
1 1 1 1
1 1 1
T T
(1.416)
The equation (1.414) can be rewritten as:
b b b
b b b b F F
C
C C
C
C C
C
C C
C
C C
C


(
(

|
|
.
|

\
|

|
|
.
|

\
|

1
1 2
,
I I I
I I I I I
I
I I I
I I I
I I I I I
I
I I I
T

1
(1.417)
In light of the equation in (1.413) and (1.417), we can draw the conclusion that:
b b
b
b b
b
b b
C
b b
b b b F F
, ,
1
,



= =
(
(

|
|
.
|

\
|

=

I I I
I I I I I
I
I I I
T
1
(1.418)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
96
which indicates that
b
, and b are coaxial tensors.
Once again, we can observe C given by the equation in (1.405). Next, we can evaluate the
derivative of the scalar-valued tensor function, ) (C = , with respect to the tensor F :
( )
( )
kl
ij
ij
kl
F
C
C

F F
F
C
C F
C
, ,
notation indicial
:
(1.419)
The derivative of tensor C with respect to F is evaluated as follows:
( )
( ) ( )
ki jl kj il
qi jl qk qj il qk
kl
qj
qi qj
kl
qi
kl
qj qi
kl
ij
F F
F F
F
F
F F
F
F
F
F F
F
C


+ =
+ =


(1.420)
Then, by substituting (1.420) into (1.419), we obtain:
( ) ( )
il
ki
lj
kj
ki jl kj il
ij
kl
C
F
C
F
F F
C

=
+

F
,
(1.421)
Due to the symmetry of C , i.e.
jl lj
C C = , we can draw the conclusion that:
( )

=
kj
jl
kj
lj
kl
F
C
F
C

2 2 ,
F C C F
F F , 2 , 2 , = =
T

(1.422)
Now, suppose that C is given by the equation
2
U U U U U = = =
T
C , where U is a
symmetric second-order tensor. To find
U ,
) (C we can use the same equation as in
(1.422), i.e.:
C C
, 2 , 2 , = = U U
U

(1.423)
Therefore, we can draw the conclusion that
C
, and U are coaxial tensors.
Let A be a symmetric second-order tensor, and ) (A = be a scalar-valued tensor
function. The following relationships hold:
T
T
T
b b b b b b
b b
b b b
b b b
b
b
b
= = + =
= =
= =
= =




and for
for
for
A
A
A
A A
A A
A
A
, ,
, 2 , 2 ,
, 2 ,
, 2 ,





(1.424)
1.7 The Voigt Notation
When dealing with symmetric tensors, it may be advantageous to just work with the
independent components. For example, a symmetric second-order tensor has 6
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
97
independent components, so, it is possible to represent these components by a column
matrix as follows:
{ }
(
(
(
(
(
(
(
(

=
(
(
(
(
(
(

=
13
23
12
33
22
11
33 23 13
23 22 12
13 12 11
T
T
T
T
T
T
T T T
T T T
T T T
T T
Voigt
ij

(1.425)
This representation is called the Voigt Notation. It is also possible to represent a second-
order tensor as:
{ }
(
(
(
(
(
(
(
(

=
(
(
(
(
(
(

=
13
23
12
33
22
11
33 23 13
23 22 12
13 12 11
2
2
2
E
E
E
E
E
E
E E E
E E E
E E E
E E
Voigt
ij

(1.426)
As we have seen before, a fourth-order tensor, C , that presents minor symmetry, i.e.
jilk ijlk jikl ijkl
C C C C = = = , has 36 6 6 = independent components. Note that, due to the
symmetry of ) (ij we have 6 independent components, and due to the symmetry of ) (kl
we have 6 independent components. In Voigt Notation we can represent these
components in a 6 -by- 6 matrix as:
[ ]
(
(
(
(
(
(
(
(

=
1313 1323 1312 1333 1322 1311
2313 2323 2312 2333 2322 2311
1213 1223 1212 1233 1222 1211
3313 3323 3312 3333 3322 3311
2213 2223 2212 2233 2222 2211
1113 1123 1112 1133 1122 1111
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C C C C C C
C
(1.427)
In addition to minor symmetry the tensor also has major symmetry, i.e.
klij ijkl
C C = , and the
number of independent components have reduced to 21. One can easily memorize the
order of the components in the matrix [ ] C if we consider the order of the second-order
tensor in Voigt Notation, i.e.:
[ ] ) 13 ( ) 23 ( ) 12 ( ) 33 ( ) 22 ( ) 11 (
) 13 (
) 23 (
) 12 (
) 33 (
) 22 (
) 11 (
(
(
(
(
(
(
(
(

(1.428)
1.7.1 The Unit Tensors in Voigt Notation
The second-order unit tensor is represented in the Voigt notation as:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
98
{ }
(
(
(
(
(
(
(
(

=
(
(
(

=
0
0
0
1
1
1

1 0 0
0 1 0
0 0 1
Voigt
ij
1 (1.429)
In the subsection 1.5.2.5.1 Unit Tensors we have defined three fourth-order unit tensors,
namely,
l l j ik ijk
= I ,
jk i ijk

l l
= I and
l l k ij ijk
= I , among which only
l l k ij ijk
= I is a
symmetric tensor. The representation of
l l k ij ijk
= I in Voigt notation can be evaluated by
observing how a symmetric fourth-order tensor is represented in (1.427), thus:
[ ]
(
(
(
(
(
(
(
(

= =
0 0 0 0 0 0
0 0 0 0 0 0
0 0 0 0 0 0
0 0 0 1 1 1
0 0 0 1 1 1
0 0 0 1 1 1
I
Voigt
k ij ijk l l
I (1.430)
where 1
11 11 1111
= = I , 1
22 11 1122
= = I , and so on.
The components of a fourth-order unit tensor,
sym
I , are represented by
( )
jk i j ik ijk

l l l
+ =
2
1
I , which in Voigt notation becomes:
[ ]
(
(
(
(
(
(
(
(

=
(
(
(
(
(
(
(
(

=
2
1
2
1
2
1
1313 1323 1312 1333 1322 1311
2313 2323 2312 2333 2322 2311
1213 1223 1212 1233 1222 1211
3313 3323 3312 3333 3322 3311
2213 2223 2212 2233 2222 2211
1113 1123 1112 1133 1122 1111
0 0 0 0 0
0 0 0 0 0
0 0 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1
I I I I I I
I I I I I I
I I I I I I
I I I I I I
I I I I I I
I I I I I I
I I
Voigt
ijkl
(1.431)
and the inverse of the equation in (1.431) becomes:
[ ]
(
(
(
(
(
(
(
(

2 0 0 0 0 0
0 2 0 0 0 0
0 0 2 0 0 0
0 0 0 1 0 0
0 0 0 0 1 0
0 0 0 0 0 1
1
I (1.432)

1.7.2 The Scalar Product in Voigt Notation
The dot product between a symmetric second-order tensor, T , and a vector n
r
, is given by
n T b
r
r
= where the components of b
r
can be evaluated as follows:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
99

+ + =
+ + =
+ + =

(
(
(

(
(
(

=
(
(
(

3 33 2 23 1 13 3
3 23 2 22 1 12 2
3 13 2 12 1 11 1
3
2
1
33 23 13
23 22 12
13 12 11
3
2
1

n T n T n T b
n T n T n T b
n T n T n T b
n
n
n
T T T
T T T
T T T
b
b
b

(1.433)
By observing how a second-order tensor is presented in Voigt notation, as in (1.425), the
scalar product (1.433) can be represented in the Voigt notation as:
[ ]

(
(
(
(
(
(
(
(

(
(
(

=
(
(
(

13
23
12
33
22
11
1 2 3
3 1 2
3 2 1
3
2
1

0 0 0
0 0 0
0 0 0
T
T
T
T
T
T
n n n
n n n
n n n
b
b
b
4 4 4 4 4 3 4 4 4 4 4 2 1
T
N
{ } [ ] { } T N
T
= b (1.434)
1.7.3 The Component Transformation Law in Voigt Notation
The component transformation law for a second-order tensor is defined as:
jl ik kl ij
a a T T =

(1.435)
or in matrix form:
T
a a a
a a a
a a a
a a a
a a a
a a a
(
(
(

(
(
(

(
(
(

=
(
(
(




33 32 31
23 22 21
13 12 11
33 23 13
23 22 12
13 12 11
33 32 31
23 22 21
13 12 11
33 23 13
23 22 12
13 12 11

T T T
T T T
T T T
T T T
T T T
T T T
(1.436)
By multiplying the matrices and by rearranging the result in Voigt notation we obtain:
{ } [ ]{ } T M T = (1.437)
where:
{ } { }
(
(
(
(
(
(
(
(

=
(
(
(
(
(
(
(
(

=
13
23
12
33
22
11
13
23
12
33
22
11
;
T
T
T
T
T
T
T
T
T
T
T
T
T T
(1.438)
and [ ] M is the transformation matrix for the second-order tensor components in Voigt
Notation. The matrix [ ] M is given by:
[ ]
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )(
(
(
(
(
(
(
(

+ + +
+ + +
+ + +
=
13 31 11 33 13 32 12 33 11 32 12 31 13 33 12 32 11 31
23 31 21 33 23 32 22 33 21 32 22 31 23 33 22 32 21 31
23 11 21 13 23 12 22 13 21 12 22 11 23 13 12 22 11 21
33 31 33 32 32 31
2
33
2
32
2
31
23 21 23 22 22 21
2
23
2
22
2
21
13 11 13 12 12 11
2
13
2
12
2
11
2 2 2
2 2 2
2 2 2
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
M

(1.439)
If the representation of tensor components is shown in (1.426), equation (1.436) in Voigt
Notation becomes:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
100
{ } [ ]{ } E N E = (1.440)
where
[ ]
( ) ( ) ( )
( ) ( ) ( )
( ) ( ) ( )(
(
(
(
(
(
(
(

+ + +
+ + +
+ + +
=
13 31 11 33 13 32 12 33 11 32 12 31 13 33 12 32 11 31
23 31 21 33 23 32 22 33 21 32 22 31 23 33 22 32 21 31
23 11 21 13 23 12 22 13 21 12 22 11 23 13 12 22 11 21
33 31 33 32 32 31
2
33
2
32
2
31
23 21 23 22 22 21
2
23
2
22
2
21
13 11 13 12 12 11
2
13
2
12
2
11
2 2 2
2 2 2
2 2 2
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
a a a a a a a a a
N

(1.441)
The matrices (1.439) and (1.441) are not orthogonal matrices, i.e. [ ] [ ]
T
M M
1
and
[ ] [ ]
T
N N
1
. However, it is possible to show that [ ] [ ]
T
N M =
1
.
1.7.4 Spectral Representation in Voigt Notation
Regarding the spectral representation of a symmetric tensor T :
A T A T

form Matricial
3
1
) ( ) (
= =

=
T
a
a a
a
n n T T (1.442)
where A is the transformation matrix between the original set and the principal space,
made up of the eigenvectors
) (

a
n . The above equation can be rewritten in terms of
components as follows:
A A A A A A
(
(
(

+
(
(
(

+
(
(
(

=
(
(
(

3
2
1
33 23 13
23 22 12
13 12 11
0 0
0 0 0
0 0 0
0 0 0
0 0
0 0 0
0 0 0
0 0 0
0 0

T
T
T
T T T
T T T
T T T
T T T

(1.443)
or
(
(
(

+
(
(
(

+
(
(
(

=
(
(
(

2
33 33 32 33 31
33 32
2
32 32 31
33 31 32 31
2
31
3
2
23 23 22 23 21
23 22
2
22 22 21
23 21 22 21
2
21
2
2
13 13 12 13 11
13 12
2
12 12 11
13 11 12 11
2
11
1
33 23 13
23 22 12
13 12 11
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
a a a a a
T
T T
T T T
T T T
T T T
(1.444)
By regarding how second-order tensors are presented in Voigt Notation as in (1.438), the
spectral representation of a second-order tensor in Voigt notation becomes:
{ }
(
(
(
(
(
(
(
(

+
(
(
(
(
(
(
(
(

+
(
(
(
(
(
(
(
(

=
(
(
(
(
(
(
(
(

=
33 31
33 32
32 31
2
33
2
32
2
31
3
23 21
23 22
22 21
2
23
2
22
2
21
2
13 11
13 12
12 11
2
13
2
12
2
11
1
13
23
12
33
22
11
a a
a a
a a
a
a
a
a a
a a
a a
a
a
a
a a
a a
a a
a
a
a
T T T
T
T
T
T
T
T
T (1.445)
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
101
1.7.5 Deviatoric Tensor Components in Voigt Notation
Observing the components of the deviatoric tensor:
(
(
(




=
) 2 (
) 2 (
) 2 (
22 11 33 3
1
23 13
23 33 11 22 3
1
12
13 12 33 22 11 3
1
T T T T T
T T T T T
T T T T T
T
dev
ij
(1.446)
dev
ij
T in Voigt notation is given by:
(
(
(
(
(
(
(
(

(
(
(
(
(
(
(
(




=
(
(
(
(
(
(
(
(

13
23
12
33
22
11
13
23
12
33
22
11
3 0 0 0 0 0
0 3 0 0 0 0
0 0 3 0 0 0
0 0 0 2 1 1
0 0 0 1 2 1
0 0 0 1 1 2
3
1
T
T
T
T
T
T
T
T
T
T
T
T
dev
dev
dev
dev
dev
dev

(1.447)
Problem 1.40: Let ) , ( t x
r
T be a symmetric second-order tensor, which is expressed in
terms of the position ) ( x
r
and time ) (t . Also, bear in mind that the tensor components,
along direction
3
x , are equal to zero, i.e. 0
33 23 13
= = = T T T .
NOTE: In the next section we will define ) , ( t x
r
T as a field tensor, i.e. the value of T
depends on position and time. As we will see later, if the tensor is independent of any one
direction at all points ) ( x
r
, e.g. if ) , ( t x
r
T is independent of the
3
x -direction, (see Figure
1.31), the problem becomes a two-dimensional problem (plane state) so that the problem is
greatly simplified.














Figure 1.31: A two-dimensional problem (2D).

a) Obtain
11
T ,
22
T ,
12
T in the new reference system (
2 1
x x ) defined in Figure 1.32.
b) Obtain the value of so that corresponds to the principal direction of T , and also
find an equation for the principal values of T .
c) Evaluate the values of
ij
T , ) 2 , 1 , ( = j i , when 1
11
= T , 2
22
= T , 4
12
= T and 45 = .
Also, obtain the principal values and principal directions.
d) Draw a graph that shows the relationship between and components
11
T ,
22
T and
12
T , and in which the angle varies from 0 to 360 .
Hint: Use the Voigt Notation, and express the results in terms of 2 .

1
x

2
x

3
x

22
T

11
T
12
T

12
T

1
x

2
x

22
T

11
T
12
T

12
T

11
T

22
T
2D
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
102









Figure 1.32: A two-dimensional problem (2D).
Solution:
a) Here we can apply the transformation law obtained in (1.437), which after removing
rows and columns associated with the
3
x -direction becomes:
(
(
(

(
(
(

+
=
(
(
(

12
22
11
21 12 22 11 12 22 11 21
22 21
2
22
2
21
12 11
2
12
2
11
12
22
11
2
2
T
T
T
T
T
T
a a a a a a a a
a a a a
a a a a
(1.448)

















Figure 1.33: Transformation law for (2D) tensor components.

The transformation matrix,
ij
a , in the plane, can be evaluated in terms of a single
parameter, :
(
(
(

=
(
(
(

=
1 0 0
0 cos sin
0 sin cos
33 32 31
23 22 21
13 12 11


a a a
a a a
a a a
a
ij
(1.449)
By substituting the matrix components
ij
a given in (1.449) into (1.448) we obtain:
(
(
(

(
(
(

=
(
(
(

12
22
11
2 2
2 2
2 2
12
22
11

sin cos cos cos sin
cos sin 2 cos sin
cos 2 sin cos
sin
sin
T
T
T
T
T
T



(1.450)
Making use of the following trigonometric identities, 2 cos 2 sin sin = ,
2 cos sin cos
2 2
= ,
2
2 cos 1
sin
2


= ,
2
2 cos 1
cos
2

+
= , (1.450) becomes:
P

22
T

11
T

1
x

2
x

12
T

11
T

12
T

22
T
P


T
A T A T =
A T A T =
T


1
x
1
x

2
x

22
T

11
T
12
T

12
T

11
T

22
T


1
x

2
x

2
x
1
x
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
103
(
(
(

(
(
(
(
(
(
(

|
.
|

\
|
|
.
|

\
|

|
.
|

\
| +
|
.
|

\
|
|
.
|

\
|
|
.
|

\
| +
=
(
(
(

12
22
11
12
22
11

2 cos
2
2
2
2
2
2
2 cos 1
2
2 cos 1
2
2
2 cos 1
2
2 cos 1
sin sin
sin
sin
T
T
T
T
T
T



Explicitly, the above components are given by:

+ |
.
|

\
|
+ |
.
|

\
|
=
|
.
|

\
| +
+ |
.
|

\
|
=
+ |
.
|

\
|
+ |
.
|

\
| +
=


2 cos
2
2
2
2
2
2
2 cos 1
2
2 cos 1
2
2
2 cos 1
2
2 cos 1
12 22 11 12
12 22 11 22
12 22 11 11
sin sin
sin
sin
T T T T
T T T T
T T T T

Reordering the previous equation, we obtain:

+ |
.
|

\
|
=
|
.
|

\
|
|
.
|

\
| +
=
+ |
.
|

\
|
+ |
.
|

\
| +
=




2 cos 2
2
2 2 cos
2 2
2 2 cos
2 2
12
22 11
12
12
22 11 22 11
22
12
22 11 22 11
11
sin
sin
sin
T
T T
T
T
T T T T
T
T
T T T T
T
(1.451)
b) Recalling that the principal directions are characterized by the lack of any tangential
components, i.e. 0 =
ij
T if j i , in order to find the principal directions in the plane, we let
0
12
= T , hence:
22 11
12
22 11
12
12
22 11
12
22 11
12
2
) 2 (
2
2 cos
2
2 cos 2
2
0 2 cos 2
2
sin
sin sin
T T
T
tg
T T
T
T
T T
T
T T
T

|
.
|

\
|
= + |
.
|

\
|
=
= =
=



Then, the angle corresponding to the principal direction is:
|
|
.
|

\
|

=
22 11
12
2
2
1
T T
T
arctg (1.452)
To find the principal values (eigenvalues) we must solve the following characteristic
equation:
( ) ( ) 0 0
2
12 22 11 22 11
2
22 12
12 11
= + + =

T T T T T T T
T T T
T T T

And by evaluating the quadratic equation we obtain:
( ) [ ] ( ) [ ] ( )
( ) [ ] ( )
4
4
2
) 1 ( 2
) 1 ( 4
2
12 22 11
2
22 11 22 11
2
12 22 11
2
22 11 22 11
) 2 , 1 (
T T T T T T T
T T T T T T T
T
+

+
=
+ +
=

By rearranging the above equation we obtain the principal values for the two-dimensional
case as:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
104
2
12
2
22 11 22 11
) 2 , 1 (
2 2
T
T T T T
T + |
.
|

\
|

+
=
(1.453)
c) We directly apply equation (1.451) to evaluate the values of the components
ij
T ,
) 2 , 1 , ( = j i , where 1
11
= T , 2
22
= T , 4
12
= T and 45 = , i.e.:

= |
.
|

\
|
=
= |
.
|

\
|
|
.
|

\
| +
=
= |
.
|

\
|
+ |
.
|

\
| +
=

+
5 . 0 90 cos 90
2
2 1
5 . 5 90 90 cos
2
2 1
2
2 1
5 . 2 90 4 90 cos
2
2 1
2
2 1
4 sin
sin 4
sin
12
22
11
T
T
T

And the angle corresponding to the principal direction is:
( ) 4375 . 41
2 1
) 4 ( 2 2
2
1
22 11
12
=


=
|
|
.
|

\
|

=
T T
T
arctg
The principal values of ) , ( t x
r
T can be evaluated as follows:

=
=
+ |
.
|

\
|

+
=
5311 . 2
5311 . 5

2 2
2
1 2
12
2
22 11 22 11
) 2 , 1 (
T
T
T
T T T T
T
d) By referring to equation in (1.451) and by varying from 0 to 360 , we can obtain
different values of
11
T ,
22
T ,
12
T , which are illustrated in the following graph:




























-6
-4
-2
0
2
4
6
8
0 50 100 150 200 250 300 350

5311 . 5
1
= T

22
T


12
T


11
T


22
T

12
T

11
T
437 . 41 =

2
T

0311 . 4
max
=
S
T

C
o
m
p
o
n
e
n
t
s


1
x
437 . 131 =

2


437 . 86 =
45

1
x

1
x
5311 . 2
2
= T

1
T

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
105
1.8 Tensor Fields
A tensor field indicates how the tensor, ) , ( t x
r
T , varies in space ( x
r
) and time ( t ). In this
section, we regard the tensor field as a differentiable function of position and time. For
more information about it, we need to define some operators, e.g. gradient, divergence, curl,
which we can use as indicators of how these fields vary in space.
A tensor field which is independent of time is called a stationary or steady-state tensor
field, i.e. ) ( x
r
T T = . However, if the field is only dependent on t then it is said to be
homogeneous or uniform. That is, ) (t T has the same value at every x
r
position.
Tensor fields can be classified according to their order as: scalar, vector, second-order
tensor fields, etc. As an example of a scalar field we can quote temperature ) , ( t T x
r
and in
Figure 1.34(a) we can see temperature distribution over time
1
t t = . Then, as an example of
a vector field we can quote velocity ) , ( t x v
r r
and Figure 1.34(b) shows velocity distribution,
in which each point is associated with a vector v
r
over time
1
t t = .












Figure 1.34: Examples of tensor fields.
Scalar Field

) , ( t x
r
= (1.454)
Vector Field
Tensorial notation
) , ( t x
r r r
v v =
Indicial notation ) , ( t
i i
x
r
v v =
(1.455)
Second-Order Tensor Field
Tensorial notation ) , ( t x
r
T T =
Indicial notation ) , ( t
ij ij
x
r
T T =
(1.456)

2
x

1
x

3
x

5
T

6
T
) , (
1
) 4 (
4
t T x
r


3
T

7
T

2
T

1
T

8
T
a) Scalar field

2
x

1
x

3
x
b) Vector field

1
t t =
1
t t = ) , (
1
t x v
r r

NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
106
1.8.1 Scalar Fields
The next analysis is carry out with reference to a stationary scalar field, i.e. ) ( x
r
= , with
continuous values of
1
/ x ,
2
/ x and
3
/ x . Then, observe that the value of the
scalar function at point ) ( x
r
is ) ( x
r
, and if we observe a second point located at ) ( x x
r r
d + ,
the total derivative (differential) of the function is defined as:


d x x x dx x dx x dx x
d d
+ + +
+
) , , ( ) , , (
) ( ) (
3 2 1 3 3 2 2 1 1
x x x
r r r
(1.457)
For any continuous function ) , , (
3 2 1
x x x , d is linearly related to
1
dx ,
2
dx ,
3
dx . This
linear relationship can be evaluated by the chain rule of differentiation as:
i i
dx d
dx
x
dx
x
dx
x
d
,
3
3
2
2
1
1

=
(1.458)
The differentiation of the components of a tensor, with respect to coordinates
i
x , is
expressed by the differential operator:
i
i
x
,



(1.459)
1.8.2 Gradient
The gradient of a scalar field
The gradient
x
r
or grad is defined as:
x
x x
r
r r
d d = (1.460)
where the operator
x
r
is known as the Nabla symbol. Expressing the equation (1.460) in
the Cartesian basis we obtain:
=

3
3
2
2
1
1
dx
x
dx
x
dx
x


[ ] [ ]
3 3 2 2 1 1 3
3
2
2
1
1

) (

) (

) (

) (

) (

) ( e e e e e e dx dx dx
x x x
+ + + + =
x x x
r r r

(1.461)
Evaluating the above scalar product we find:
3
3
2
2
1
1
3
3
2
2
1
1
) ( ) ( ) ( dx dx dx dx
x
dx
x
dx
x
x x x


x x x
r r r
+ + =


(1.462)
Therefore, we can draw the conclusion that the
x
r
components in the Cartesian basis
are:
3
3
2
2
1
1
) ( ; ) ( ; ) (
x x x

x x x
r r r

(1.463)
Hence, the gradient in terms of components is defined as such:

3
3
2
2
1
1

e e e
x x x

x
r

(1.464)
The Nabla symbol
x
r
is defined as:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
107
i i i
i
x
e e

,

=
x
r

Nabla symbol (1.465)


The geometric meaning of
x
r

The direction of
x
r
is normal to the equiscalar surface, i.e. it is perpendicular to
the isosurface const = . The direction of
x
r
points to the direction where is
increasing the most, (see Figure 1.35).
The magnitude of
x
r
is the rate of change of , i.e. the gradient of .
The normal vector to this surface is obtained as follows:

x
x
r
r

= n


(1.466)
The surface const = , called the surface level, or isosurface or equiscalar surface, is the
surface formed by points which all have the same value of , so, if we move along the
level surface the values of the function do not change.
The gradient of a vector field ) ( x
r r
v :
v v
r r
r
x
) ( grad (1.467)
Using the definition of
x
r
, given in (1.465), the gradient of the vector field becomes:
j i j i j j i i j
j
i i
x
e e e e e
e
v

)

(
, ,
= =

= v v
v r
r
x

(1.468)










Figure 1.35: Gradient of .
Therefore, we can define the gradient of a tensor field )) , ( ( t x
r
in the Cartesian basis as:
j
j
x
e

) (
) (


=
x
r

Gradient of a tensor field in the
Cartesian basis
(1.469)
As noted, the gradient of a vector field becomes a second-order tensor field, whose
components are:

1
c const = =

2
c const = =

3
c const = =

x
r

n



3 2 1
c c c > >
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
108
(
(
(
(
(
(
(

3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
,
x x x
x x x
x x x
x
j
i
j i
v v v
v v v
v v v
v
v (1.470)
The gradient of a second-order tensor field ) ( x
r
T :
k j i k ij k
k
j i ij
T
x
T
e e e e
e e
T

)

(
,
=


=
x
r
(1.471)
and its components are represented by:
( )
k ij ijk
T
,
T
x
r

(1.472)
Problem 1.41: Find the gradient of the function
2 1
) cos( ) , (
1 2 1
x x
x x x f exp + = at the point
) 1 , 0 (
2 1
= = x x .
Solution: By definition, the gradient of a scalar function is given by:
2
2
1
1

e e
x
f
x
f
f

=
x
r

where:
2 1
2 1
1
) sin(
x x
x x
x
f
exp + =

;
2 1
1
2
x x
x
x
f
exp =


[ ] [ ]
2 1 1 2 1 2 1

) sin( ) , (
2 1 2 1
e e
x x x x
x x x x x f exp exp + + =
x
r
[ ] [ ]
1 2 1

2

1 ) 1 , 0 ( e e e = + = f
x
r

Problem 1.42: Let ) ( x
r r
u be a stationary vector field. a) Obtain the components of the
differential u
r
d . b) Now, consider that ) ( x
r r
u represents a displacement field, and is
independent of
3
x . With these conditions, graphically illustrate the displacement field in
the differential area element
2 1
dx dx .
Solution: According to the differential and gradient definitions, it holds that:











Thus, the components are defined as:
(
(
(

(
(
(
(
(
(
(

=
(
(
(

=
3
2
1
3
3
2
3
1
3
3
2
2
2
1
2
3
1
2
1
1
1
3
2
1

dx
dx
dx
x x x
x x x
x x x
d
d
d
dx
x
d
j
j
i
i
u u u
u u u
u u u
u
u
u
u
u
) ( x
r r
u
x
r
d

1
x

2
x

3
x
) ( x x
r r r
d + u
x
r

x x
r r
d +

x
x x x
x
r r r
r r r r r r
r
d d
d d
=
+
u u
u u u

) ( ) (

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
109
or:

=
3
3
3
2
2
3
1
1
3
3
3
3
2
2
2
2
1
1
2
2
3
3
1
2
2
1
1
1
1
1
dx
x
dx
x
dx
x
d
dx
x
dx
x
dx
x
d
dx
x
dx
x
dx
x
d
u u u
u
u u u
u
u u u
u

with

+ + + =
+ + + =
+ + + =
) , , ( ) , , (
) , , ( ) , , (
) , , ( ) , , (
3 2 1 3 3 3 2 2 1 1 3 3
3 2 1 2 3 3 2 2 1 1 2 2
3 2 1 1 3 3 2 2 1 1 1 1
x x x dx x dx x dx x d
x x x dx x dx x dx x d
x x x dx x dx x dx x d
u u u
u u u
u u u

As the field is independent of
3
x , the displacement field in the differential area element is
defined as:

= + + =

= + + =
2
2
2
1
1
2
2 1 2 2 2 1 1 2 2
2
2
1
1
1
1
2 1 1 2 2 1 1 1 1
) , ( ) , (
) , ( ) , (
dx
x
dx
x
x x dx x dx x d
dx
x
dx
x
x x dx x dx x d
u u
u u u
u u
u u u



or:

+ = + +

+ = + +
2
2
2
1
1
2
2 1 2 2 2 1 1 2
2
2
1
1
1
1
2 1 1 2 2 1 1 1
) , ( ) , (
) , ( ) , (
dx
x
dx
x
x x dx x dx x
dx
x
dx
x
x x dx x dx x
u u
u u
u u
u u

Note that the above equation is equivalent to the Taylor series expansion taking into
account only up to linear terms. The representation of the displacement field in the
differential area element is shown in Figure 1.36.




















NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
110















































Figure 1.36: Displacement field in the differential area element.



2
2
1
dx
x
u


1
1
2
dx
x
u


2
dx

2
u
A
O

1
dx
B
A

2
dx
B

1
u

1
1
1
1
dx
x

+
u
u
O

1
dx

1 1
, u x

2 2
, u x
B
A
A

2
2
2
2
dx
x

+
u
u
B
O
+
) (
1
u
) , (
2 1
x x ) , (
2 1 1
x dx x +
) , (
2 2 1
dx x x +
) , (
2 2 1 1
dx x dx x + +

1
x

2
x
u
r
d
) (
2
u

2
2
2
1
1
2
2
dx
x
dx
x

+
u u
u

2
2
1
1
1
1
1
dx
x
dx
x

+
u u
u

2
2
2
2
dx
x

+
u
u

2
2
1
1
dx
x

+
u
u

1
1
1
1
dx
x

+
u
u

1
1
2
2
dx
x

+
u
u

1
dx

2
dx

4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 3 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 2 1


4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 8 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 4 7 6

=
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
111
1.8.3 Divergence
The divergence of a vector field, ) ( x
r r
v , is denoted as follows:
v v
r r
r

x
) ( div (1.473)
which by definition is:
) ( ) ( v 1 v v v
r r r r
r r r
x x x
Tr div = = : (1.474)
Then:
[ ] [ ]
3
3
2
2
1
1
, , ,

x x x
k k jl ik kl j i l k kl j i j i

=
= = = =
v v v
v v v e e e e 1 v v : :
r r
r r
x x

(1.475)
or
[ ] [ ]
[ ]
[ ]
[ ]
k
k
i i
k i k i
k i lj kl j i
l k kl j i j i
x
e
e
e e
e e
e e e e 1 v v




,
,
,

=
=
=
= =
v
v
v
v

: :
r r
r r
x x

(1.476)
Which we can use to insert the following operator into the Cartesian basis:
k
k
x
e

) (
) (


=
x
r

Divergence of ) ( in Cartesian
basis
(1.477)
We can also verify that, when divergence is applied to a tensor field its rank decreases by
one order.
Divergence of a second-order tensor field ) ( x
r
T
The divergence of a second-order tensor field T is denoted by 1 T T :
x x
r r
= , which
becomes a vector:
i k ik
i jk
k
ij
k
k
j i ij
x
T
x
T
e
e
e
e e
T T

)

(
,
T
div
=

=


=

x
r


(1.478)
NOTE: In this text book, when dealing with gradient or divergence of a tensor field, e.g. v
r
r
x

(the gradient of the vector field), T
x
r
(the gradient of a second-order tensor field), T
x
r

(divergence of a second-order tensor field), this does not indicate that we are making a
tensor operation between a vector and a tensor, i.e. ) ( ) ( v v
r
r
r
r r

x x
, ) ( ) ( T T
x x
r r
r

and ) ( ) ( T T
x x
r r
r
and so on. In this textbook,
x
r
is an operator which must be
applied to the entire tensor field, so, the tensor must be inside the operator, (see equations
(1.477) and (1.469)). Nevertheless, it is possible to relate v
r
r
x
, T
x
r
or T
x
r
to tensor
operations between tensors, and it is easy to show that:
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
112

) ( ) ( ) ( ) (
) ( ) (
) ( ) (
T
T T T
T T
v v
= =
=
=
x x x
x x
x x
r r r
r r
r r
r r
r
r
r r





(1.479)
Once the Nabla symbol is defined we introduce the Laplacian operator
2
as:
kk k k
i i
ij
j i
i j
j i
x x x
x x x x x x
, , ,

2
3
2
2
2
2
2
1
2
2
2
2
= =

=
|
|
.
|

\
|
|
|
.
|

\
|

= =
x
x x x
r
r r r

e e

(1.480)
Then, the vector Laplacian of a vector field, ) ( x
r r
v , is given by:
[ ] [ ]
kk i
i i
components
,
2 2
) ( ) ( v = = = v v v v
r r r r
r r r r r r
x x x x x x
(1.481)
Problem 1.43: Let a
r
and b
r
be vectors. Show that the following identity
b a b a
r
r
r
r
r r r
+ = +
x x x
) ( holds.
Solution:
Observing that
j j
e a

a =
r
,
k k
e b

b =
r
,
i
i
x

= e

x
r
, we can express ) ( b a
r
r
r
+
x
as:
b a e e e e e
e e
r
r
r r
+ =

+
x x

i
i
i
i
i k
i
k
i j
i
j
i
i
k k j j
x x x x x
b a b
a b a

)

(

Working directly with indicial notation we obtain:
b a b a
r
r
r
r
r r r
+ = + = + = +
x x x

i i i i i i i
, , ), ( ) ( b a b a
Problem 1.44: Find the components of b a
r
r
r
) (
x
.
Solution: Bearing in mind that
j j
e a

a =
r
,
k k
e b

b =
r
,
i
i
x

= e

x
r
( 3 , 2 , 1 = i ), the following is
true:
j
k
j
k j
i
j
ik k k k i j
i
j
k k i
i
j j
x x x x
e e e e e e e
e
b a

)

(

)

(
) (

=
|
|
.
|

\
|

=
|
|
.
|

\
|

=
a
b
a
b b
a
b
a

r
r
r
x

Expanding the dummy index k , we obtain:
3
3
2
2
1
1
x x x x
j j j
k
j
k

a
b
a
b
a
b
a
b
Thus,
3
3
3
2
3
2
1
3
1
3
2
3
2
2
2
1
2
1
3
1
3
2
1
2
1
1
1
3
2
1
x x x
j
x x x
j
x x x
j

=
a
b
a
b
a
b
a
b
a
b
a
b
a
b
a
b
a
b

Problem 1.45: Prove that the following relationship is valid:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
113
T
T T T
x x x
r r r
r r
r
=
|
|
.
|

\
|
q q
q
2
1 1

where ) , ( t x
r
r
q is an arbitrary vector field, and ) , ( t T x
r
is a scalar field.
Solution:
(scalar)
1 1
1 1
2
,
2
,
,
T
T T
T
T T T T x T
i i i i
i
i i
i
x x
x
r r
r
r r
r

=
= |
.
|

\
|
|
.
|

\
|

=
|
|
.
|

\
|
q q
q
q q
q q

1.8.4 The Curl
The curl of a vector field
The curl (or rotor) of a vector field, ) ( x
r r
v is denoted by v v v
r
r
r r
r

x
) ( ) ( rot curl , and is
defined in the Cartesian basis as:
) (

) (

=
j
j
x
e
x
r
r

The curl (rotor) of a tensor field in
the Cartesian basis
(1.482)
Note that the curl is already a tensor operator between two vectors. Using the definition of
the vector product we obtain the curl of a vector field as:
i j k ijk i ijk
j
k
k j
j
k
k k j
j
x x x
e e e e e e v v

)

) (
,
v
v v
v rot =

= =
r
r
r
r
x

(1.483)
where
ijk
is the permutation symbol defined in (1.55). Moreover, we have applied the
definition
i ijk k j
e e e

= and we can also note that:
3
2
1
1
2
2
1
3
3
1
1
3
2
2
3
,
3 2 1
3 2 1
3 2 1


) (
e e e
e
e e e
v v
|
|
.
|

\
|

+
|
|
.
|

\
|

+
|
|
.
|

\
|

=
=

= =
x x x x x x
x x x
i j k ijk
v v v v v v
v
v v v
rot
r
r
r
r
x

(1.484)
We can verify that the antisymmetric part of a vector field gradient, which is illustrated by
W v
skew
) (
r
r
x
, has as components:
[ ]
(
(
(

=
(
(
(


=
(
(
(

=
(
(
(
(
(
(
(
(

|
|
.
|

\
|

|
|
.
|

\
|

|
|
.
|

\
|

|
|
.
|

\
|

|
|
.
|

\
|

|
|
.
|

\
|

=
0
0
0
0
0
0
0
0
0
0
2
1
2
1
2
1
0
2
1
2
1
2
1
0
) (
1 2
1 3
2 3
23 13
23 12
13 12
32 31
23 21
13 12
3
2
2
3
3
1
1
3
2
3
3
2
2
1
1
2
1
3
3
1
1
2
2
1
,
w w
w w
w w
x x x x
x x x x
x x x x
skew
j i ij
skew
W W
W W
W W
W W
W W
W W
v v v v
v v v v
v v v v
v v
r
r
x

(1.485)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
114
where
1
w ,
2
w ,
3
w are the components of the axial vector w
r
associated with W, (see
subsection: 1.5.2.2.2. Antisymmetric Tensor).
With reference to the definition of the curl in (1.484) and the relationship in (1.485), we can
conclude that:
( )
w
x
r
r
r
r
r
2

2

2

2

) (
3 3 2 2 1 1
3 21 2 13 1 32
3
2
1
1
2
2
1
3
3
1
1
3
2
2
3
=
+ + =
+ + =
|
|
.
|

\
|

+
|
|
.
|

\
|

+
|
|
.
|

\
|

=
e e e
e e e
e e e v v
w w w
x x x x x x
W W W
v v v v v v
rot
(1.486)
And, if we use the identity in (1.141), we obtain:
( ) v v v v W
r r
r
r r r
r
= =
x
w
2
1
(1.487)
It could be interesting to note that the equation in (1.486) can be obtained by means of
Problem 1.18, in which we showed that ) (
2
1
x a
r r
is the axial vector associated with the
antisymmetric tensor
skew
) ( a x
r r
. Therefore, the axial vector associated with the
antisymmetric tensor [ ]
skew
skew
) ( ) ( ) (
r
r r
r
= = v v W
x
is the vector ( ) v
r
r
r

x

2
1
.
As we can see, the curl describes the rotational tendency of the vector field.
Summary
Divergence

x
r
) ( div
Gradient

x
r
) ( grad
Curl

x
r
r
) ( rot
Scalar vector
Vector Scalar Second-order tensor Vector
Second-order tensor Vector Third-order tensor Second-order tensor

We can now present some equations:
) ( ) ( ) ( ) ( a a a a
r r
r
r
r
r
r r r
+ = =
x x x
rot
The result of the algebraic operation ) ( a
r
r
r

x
is a vector, whose components are
given by:
[ ]
i i
k j ijk i
k j ijk j k ijk
j k k j ijk
j k ijk i
) ( ) (
) ( ) (
) (
) ( ) (
, ,
, ,
,
a a
a
a
r r
r
r
r
r r
r r
r
=
=
=
+ =
=
x x
x x
x

a
a a
a a
a

(1.488)
We can use the above equation to check that the relationship
) ( ) ( ) ( ) ( a a a a
r r
r
r
r
r
r r r
+ = =
x x x
rot holds.

1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
115
a b b a b a a b b a
r
r r
r
r
r r
r r
r
r
r r r r r
+ = ) ( ) ( ) ( ) ( ) (
x x x x x
(1.489)
The components of the vector product ) ( b a
r
r
are given by
j i kij k
b a = ) ( b a
r
r
, thus:
[ ] ) ( ) ) (
, , ,
(
p j i j p i lpk kij p j i kij lpk l
b a b a b a + = = b a
r
r
r
r
x
(1.490)
Regarding that
ijk kij
= and
jl ip jp il lpk ijk
= , the above equation becomes:
[ ]
p l p p p l l p p p p l
p j i jl ip p j i jp il j p i jl ip j p i jp il
p j i j p i jl ip jp il p j i j p i lpk kij l
, , , ,
, , , ,
, , , ,
) )( ( ) ( ) (
b a b a b a b a
b a b a b a b a
b a b a b a b a
+ =
+ =
+ = + =

b a
r
r
r
r
x


(1.491)
We can also verify that [ ]
p p l l
b a
,
) ( = b a
r
r
r
x
, [ ]
l p p l
b a
,
) ( = b a
r
r
r
x
, [ ]
p p l l ,
) ( b a = a b
r
r
r
x
,
[ ]
p l p l ,
) ( b a = a b
r
r
r
x
.
a a a
r r r
r r
r r r r r
2
) ( ) (
x x x x x
= (1.492)
The components of ) ( a
r
r
r

x
are given by
3 2 1
r
r
r
i
j k ijk i
c
a
,
) ( = a
x
, thus:
[ ]
jl k ijk qli l j k ijk qli l i qli q , , , ,
) ) ( ( a a c = = = a
r
r r
r r
x x

(1.493)
Once again considering that
lj qk lk qj jki qli ijk qli
= = , the above equation becomes:
[ ]
ll q kq k
jl k lj qk jl k lk qj jl k lj qk lk qj jl k ijk qli q
, ,
, , , ,
) ( ) (
a a
a a a a
=
= = = a
r
r r
r r
x x

(1.494)
where [ ]
kq k q ,
) ( a = a
r
r r
x x
and [ ]
ll q q ,
2
a = a
r
r
x
.
) ( ) ( ) (
2

x x x x x
r r r r r
+ = (1.495)
) ( ) ( ) ( ) (
2
, , , , ,

x x x x x
r r r r r
+ = + = =
i i ii i i

(1.496)
where and are scalar fields. Other interesting equations derived from the above are:
) ( ) ( ) (
) ( ) ( ) (
2
2


x x x x x
x x x x x
r r r r r
r r r r r




+ =
+ =
(1.497)
After subtracting the above two identities we obtain:


2 2
2 2
) (
) ( ) (
x x x x x
x x x x x x
r r r r r
r r r r r r


=
=


(1.498)
1.8.5 The Conservative Field
A vector field, ) , ( t x
r
r
b , is said to be conservative if there exists a differentiable scalar field,
) , ( t x
r
, so that:

x
r
r
= b (1.499)
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
116
If the function satisfies the relation (1.499), then is a potential function of ) , ( t x
r
r
b .
A necessary but insufficient condition for ) , ( t x
r
r
b to be conservative is that 0 b
r r
r
r
=
x
. In
other words, given a conservative field, the curl b
r
r
r

x
equals zero. However, if the curl
of a vector field equals zero, this does not necessarily mean that the field is conservative.
Problem 1.46: Let be a scalar field, and u
r
be a vector field. a) Show that
( ) 0 = v
x x
r
r
r r
and ( ) 0
r r
r r
=
x x
.
b) Show that ( ) [ ] [ ] ) ( ) ( ) ( ) )( ( v v v v v v v v
x x x x x x x x
r
r
r r r
r
r
r
r r r
r r
r r r r r r r r
+ = ;
c) Referring v
x
r r
r
= , show that
r r
r
r
r
r r r r r
2 2 2
) ( ) (
x x x x x
v v = = .

Solution:
Regarding that:
i j k ijk
v e

,
= v
x
r
r
r

( ) ( ) ( ) ( )
ji k ijk j k
i
ijk il j k
l
ijk l i j k ijk
l
v v
x
v
x
v
x
, , , ,


=

= e e v
x x
r
r
r r

The second derivative of v
r
is symmetrical with ij , i.e.
ij k ji k
v v
, ,
= , while
ijk
is
antisymmetric with ij , i.e.,
jik ijk
= , thus:
0
, 3 3 , 2 2 , 1 1 ,
= + + =
ji ij ji ij ji ij ji k ijk
v v v v
We can observe that
ji ij
v
, 1 1
equals the double scalar product by using a symmetric and an
antisymmetric tensor, so 0
, 1 1
=
ji ij
v .
Likewise, we can show that:
( ) 0 e e
r r
r r
= = =
i i i kj ijk

, 0
x x

b) Denoting by v
x
r
r
r
r
= we obtain:
( ) [ ] ) ( v v v
x x x
r r
r
r r
r r
r r r
=
Observing the equation in (1.489), it holds that:

r r r r r r r r r r
r
r r r r r
+ = ) ( ) ( ) ( ) ( ) ( v v v v v
x x x x x

Note that 0 ) ( = = v
x x x
r
r
r
r r r
. Then, we can draw the conclusion that:
[ ] ) ( ) ( ) ( ) )( (
) ( ) ( ) ( ) (
v v v v v v
v v v v
x x x x x x
x x x x
r
r
r r r
r
r
r
r
r r r r r r r r
r
r r r r r r
r r r r
+ =
+ =





c) Observing the equation in (1.492) we obtain:

r
r
r
r
r r
r r
r r r
r r r r r
=
=

x x x
x x x x x
v
v v v


) (
) ( ) (
2

Applying the curl to the above equation we obtain:
[ ] ) ( ) ( ) (
2

0
r
r r
4 4 4 3 4 4 4 2 1
r
r
r
r
r r r r r r r
r
=
=

x x x x x x x
v v
Referring once again to the equation in (1.492) to express the term ) (
r
r r
r r

x x
:
[ ]
) (
) ( ) ( ) ( ) (
2
2 2 2
0
v
v v
x x
x x x x x x x x x x x
r
r
r
4 4 3 4 4 2 1
r
r
r r r
r r
r
r
r r
r r r r r r r r r r r
=
+ = + = =
=




1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
117
1.9 Theorems Involving Integrals
1.9.1 Integration by Parts
Integration by parts states that:

=
b
) ( ) ( ) ( ) ( ) ( ) (
a
b
a
b
a
dx x u x v x v x u dx x v x u (1.500)
where
dx
dv
x v = ) ( , and the functions ) (x u , ) (x v are differentiable in b x a .
1.9.2 The Divergence Theorem
Given a domain B with a volume V , and bounded by the surface S , (see Figure 1.37), the
divergence theorem, also called the Gauss theorem, applied to the vector field states that:


= =
= =
S
i i
S
i i
V
i i
S S V
dS dS dV
d dS dV


,
v n v v
S
x
r
r r r
r
v n v v
(1.501)
where n

is the outward unit normal to surface S .













Figure 1.37.
Let T be a second-order tensor field defined in the domain B. The divergence theorem
applied to this field is defined as:

= =
S S V
d dS dV S
x
r
r

T n T T

= =
S
j ij
S
j ij
V
j ij
dS dS dV

,
T n T T (1.502)
By using the divergence theorem we can also demonstrate that:
n


B
S
dS
x
r


1
x

2
x

3
x
S
r
d
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
118
[ ]



= =
= + =
= =
S
j k
V
j k
S
j k
V
j i ik i j ik
S
j i ik
V
j i ik
V
j k
dS x dV x
dS x dV x x
dS x dV x dV x

), ( ), (
,
, ,
n
n
n


(1.503)
in which we have assumed that
ikj j ik
0
,
= . Additionally, by observing that
kj j k
x =
,
, we
can obtain:


=
= =
S
S
j k kj
S
j k
V
kj
dS V
dS x V dS x dV

n 1 x
r
n n
(1.504)
Given a second-order tensor defined in the domain B, the following is valid:
[ ]
[ ]



= + =
= + =
= =
S
k jk i
V
k jk i jk ik
S
k jk i
V
k jk i jk k i
S
k jk i
V
k jk i
V
k jk i
dS x dV x
dS x dV x x
dS x dV x dV x

), ( ), (
,
, ,
n
n
n


(1.505)
Hence proving that:


=
=

V
T
S V
V
ji
S
k jk i
V
k jk i
dV dS dV
dV dS x dV x
)

,
n x x
x
r r
r

n
(1.506)
or
( )

=
V
T
S V
dV d dV S x x
x
r
r r
r
(1.507)
Likewise, one can prove that:

=
V S V
dV dS dV )

( ) ( n x x
x
r r
r
(1.508)
Problem 1.47: Let be a domain bounded by as shown in Figure 1.38. Further
consider that m is a second-order tensor field and is a scalar field. Show that the
following relationship holds:
[ ] [ ] [ ]

=

d d d
x x x x x
r r r r r
) (

) ( ) ( m n m m:
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
119
[ ] [ ]

=

d d d
i j ij j ij i ij ij
,

) , ( ,
,
m n m m











Figure 1.38

Solution: We could directly apply the definition of integration by parts to demonstrate the
above relationship. But, here we will start with the definition of the divergence theorem.
That is, given a tensor field v
r
, it is true that:

= =

d d d d
j j j j
indicial



,
n v v
x
n v v
r r
r

Observing that the tensor v
r
is the result of the algebraic operation m v =
x
r
r
and the
equivalent in indicial notation to
ij i j
m v , = , and by substituting it in the above equation
we obtain:
[ ]
[ ]
[ ] [ ]



=
= +
= =






d d d
d d
d d d d
j ij i j ij i ij ij
j ij i j ij i ij ij
j ij i
j
ij i j j j j
,

, ,

, , ,

, ,

,
,
,
,
m n m m
n m m m
n m m n v v

The above equation in tensorial notation becomes:
[ ] [ ] [ ]

=

d d d ) (

) ( ) ( m n m m
x x x x x
r r r r r
:
NOTE: Consider now the domain defined by the volume V , which is bounded by the
surface S with the outward unit normal to the surface n

. If N
r
is a vector field and T is a
scalar field, it is also true that:


=
=
V S V
V
i j i
S
j i i
V
ij i
dV T dS T dV T
dV T N dS T N dV T N

) ( ) (
,

, ,
,
N N N
x x x x x
r r r
r r r r r
n
n

where we have directly applied the definition of integration by parts.

n


1
x

2
x
NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
120
1.9.3 Independence of Path
A curve which connects two points A and B is denoted by the path from A to B , (see
Figure 1.39). We can then establish the condition by which a line integral is independent of
path, (see Figure 1.39).








Figure 1.39: Path independence.
Let ) ( x
r
r
b be a continuous vector fields, then the integral


C
r
r
r
d b is independent of the path if
and only if b
r
is a conservative field. This means that there is a scalar field so that
x
r
r
= b .
Regarding the above, we can draw the conclusion that:




|
|
.
|

\
|

= + +
=
B
A
B
A
B
A
x
B
A
r r
r r
r r
r r
r
r
d
x x x
d
d d
3
3
2
2
1
1
3 3 2 2 1 1

)

( e e e e e e
b

b b b


(1.509)
Thus
3
3
2
2
1
1
; ;
x x x

=

b b b
(1.510)

As the field is conservative, the curl of b
r
is the zero vector:
i
x x x
0

3 2 1
3 2 1
3 2 1
=

=
b b b
e e e
0 b
r r r
r
x
(1.511)
We can therefore conclude that:

2
1
1
2
1
3
3
1
3
2
2
3
2
1
1
2
1
3
3
1
3
2
2
3
0
0
0
x x
x x
x x
x x
x x
x x
b b
b b
b b
b b
b b
b b
(1.512)
Therefore, if the above condition is not satisfied, the field is not conservative.

1
x

2
x
3
x
A
B
1
C
r
r
d
b
r


2
C
If

=
2 1
C C
r r
r
r
r
r
d d b b
b
r
- Conservative field
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
121
1.9.4 The Kelvin-Stokes Theorem
Let S be a regular surface, (see Figure 1.40), and ) , ( t x
r
r
F be a vector field. According to the
Kelvin-Stokes Theorem:

= =

dS d d n F F F

) ( ) (
r r r r r r
r r
r
x x
S
(1.513)
If p

denotes the unit vector tangent to the boundary , the Stokes theorem becomes:

= =

dS d d n F F p F

) ( ) (

r r r r r r
r r
x x
S
(1.514)
With reference to the vector representation in the Cartesian basis:
3 3 2 2 1 1

e e e F F F F + + =
r
,
3 3 2 2 1 1

e e e dS dS dS d + + = S
r
,
3 3 2 2 1 1

e e e dx dx dx d + + =
r
, the components of the curl of F
r

are given by:
( )
3
2
1
1
2
2
1
3
3
1
1
3
2
2
3
3 2 1
3 2 1
3 2 1


e e e
e e e
F
|
|
.
|

\
|

+
|
|
.
|

\
|

+
|
|
.
|

\
|

=
x x x x x x x x x
i
F F F F F F
F F F
r r
r
x
(1.515)
Next, the Stokes theorem expressed in terms of components becomes:
3
2
1
1
2
2
1
3
3
1
1
3
2
2
3
3 3 2 2 1 1
dS
x x
dS
x x
dS
x x
dx dx dx
|
|
.
|

\
|

+
|
|
.
|

\
|

+
|
|
.
|

\
|

= + +

F F F F F F
F F F
(1.516)









Figure 1.40: Stokes theorem.
In the special case when the surface S coincides with the plane , (see Figure 1.41), the
equation (1.516) remains valid. Then, if the domain coincides with the plane
2 1
x x ,
the equation (1.513) becomes:

=

dS d
3

) ( e F F
r r r
r
r
x

(1.517)
which is known as the Stokes theorem in the plane or Greens theorem, which is expressed in
terms of components as:
3
2
1
1
2
2 2 1 1
dS
x x
dx dx
|
|
.
|

\
|

= +

F F
F F
(1.518)

2
x

3
x

1
x
n


S


p


NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
122












Figure 1.41.











Figure 1.42: Greens theorem.
1.9.5 Greens Identities
Let F
r
be a vector field, and by applying the divergence theorem we obtain:

=
S V
dS dV

n F F
r r
r
x
(1.519)
With reference to the equations (1.496) and (1.498), i.e.:
) ( ) ( ) (
2

x x x x x
r r r r r
+ = (1.520)

2 2
) (
x x x x x
r r r r r
= (1.521)
and also regarding that
x
r
r
= F , and by substituting (1.520) into (1.519) we obtain:



2
x

1
x

3
x
n



2
x

1
x

3
x


3

e dS d = S
r



3

e
1 TENSORS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
123


=
= +


V S V
S V
dV dS dV
dS dV

) ( ) (

) ( ) (
2
2


x x x x
x x x x
r r r r
r r r r


n
n
(1.522)
which is known as Greens first identity.
Now, if we substituting (1.521) into (1.519) we obtain:

=
S V
dS dV

) (
2 2
n
x x x x
r r r r
(1.523)
which is known as Greens second identity.
Problem 1.48: Let b
r
be a vector field, which is defined as v b
r
r r
r
=
x
. Show that:

=
V
i i
S
i i
dV S d ,

b n b
where ) ( x
r
= .

Solution: The Cartesian components of v b
r
r r
r
=
x
are
j k ijk i ,
v b = and by substituting
them in the above surface integral we obtain:

=
S
i j k ijk
S
i i
dS dS

,
n v n b
Applying the divergence theorem we obtain:


= + =
+ =
= =
V V
i i ji k ijk j k ijk i
V
ji k ijk j k i ijk
V
i j k ijk
S
i j k ijk
S
i i
dV dV
dV
dV dS dS
i
, ) , (
) , (
), (

0
, ,
, ,
, ,
b v v
v v
v n v n b
b
43 42 1 43 42 1















NOTES ON CONTINUUM MECHANICS
Notes on Continuum Mechanics(Springer/CIMNE) - (Chapter 01) - By: Eduardo W.V. Chaves
124

Das könnte Ihnen auch gefallen