You are on page 1of 70

Chapter

KINEMATICS OF CONTINUA

1.1. Material and Spatial Descriptions of Continuum Motion


1.1.1. Lagrangian and Eulerian Coordinates. The Motion Law.
consider a continuum

B.

correspondence between every material point

= OM

t=0
M B and

Due to Axiom 2, at time

Let us

there is a one-to-one
its radius-vector

x=

O
ei . Denote Cartesian coordinates of

i ) and introduce curvilinear coordinates X i of the


xi (x = xi e
point M in the form of some differentiable one-to-one functions

in a Cartesian coordinate system

the radius-vector by
same material

xi = xi (X k ).
Since

i ,
x=xe

(1.1)

the relationship (1.1) takes the form

x = x(X k ).
Let us fix curvilinear coordinates of the point
of the continuum

(1.2)
M,

and then material points

are considered to be numbered by these coordinates

X i.

i
For any motion of the continuum B , coordinates X of material points are
considered to remain unchanged; they are said to be 'frozen' into the medium
and move together with the continuum. Coordinates
for a material point

Xi

introduced in this way

are called Lagrangian (or material).

Due to Axiom 3, at every time

there is a one-to-one correspondence be-

X i and its radius-vector


i
i
with Cartesian coordinates x , where x and x depend on t. This
i
i
means that there is a connection between Lagrangian X , and the Cartesian x
coordinates of point M and time, i.e. there exist functions in the form (0.3)
tween every point

x = OM

MB

with Lagrangian coordinates

xi = xi (X k , t)

X k VX .

These functions determine a motion of the material point


coordinate system

E3a .
continuum B .

O
ei

of the motion of the

of space

(1.3)
M

in the Cartesian

The relationships (1.3) are said to be the law

xi in (1.3) are called Eulerian (or spatial) coordinates of


material point M.
i and the coordinate system O
ei is the same for all times t,
Since x = x e
i
Coordinates

the
the

equivalent form of the motion law follows from (1.2):

x = x(X k , t).

(1.4)

Chapter 1. Kinematics of Continua

Since the consistency conditions (0.4) must be satisfied, from (1.2) and (1.4)
we get the relationships

x(X k , 0) = x(X k ),
Here the initial time
time

t=0

Figure 1.1.

t=0

xi (X k , 0) = xi (X k ).

is considered as the time

we introduced Lagrangian coordinates

Xi

t1

in (0.4), because just at

of point

The motion of a continuum: positions of continuum

(1.5)

M.

and material point

in

reference and actual configurations

Unless otherwise stipulated, functions (1.3) are assumed to be regular in the


domain

VX R3

for all

t,

thus there exist the inverse functions

X k = X k (xi , t)

xi Vx R3 .

V = W(B , 0) in a fixed coordinate system O


ei , which
continuum B at the initial time t = 0, is called the reference

The closed domain


is occupied by

configuration

at the time

K, and the domain V = W(B, t) occupied by


t > 0 is called the actual configuration K.

the same continuum

Figure 1.1 shows a geometric picture of the motion of a continuum from the

reference configuration

to the actual one

at time

in space

E3a .

It should be noticed that if the continuum motion law (1.3) (or (1.4)) is known,
then one of the main problems of continuum mechanics (to determine coordinates
of all material points of the continuum at
any time) will be resolved. However, in actual problems of continuum mechanics this
law, as a rule, is unknown and must be
found by solving some mathematical problems, whose statements are to be formulated. One of our objectives is to derive
these statements.
Figure 1.2. Extension of a beam

Example 1.1. Let us consider a continuum

B,

which at time

t=0

in the reference con-

1.1. Material and Spatial Descriptions

figuration

h3 ,

K is a rectangular parallelepiped (a beam) with edge lengths h1 , h2

and in an actual configuration

t>0

at

and

the continuum is also a rectangular

h1 , h2

parallelepiped but with different edge lengths:

and

h3 .

We assume that

corresponding sides of both the parallelepipeds lie on parallel planes, and for

(x2 , x3 ),

one of the sides, which for example is situated on the plane


of diagonals' intersection in

and in

points

are coincident (Figure 1.2). Then the

motion law (1.3) for this continuum takes the form

x = k (t) X ,
i.e. coordinates
and

xi , xi = X i

k (t) = h (t)/h

in

and

are proportional,

K,

In

of any material point

(1.6)

2, 3,

is the proportion function. The motion law (1.6) is called

the beam extension law.

Example 1.2.

= 1,

let a continuum

be a

rectangular parallelepiped oriented as shown in


Figure 1.3; its motion law (1.3) has the form

1
1
2
x = X + a(t)X ,
2
2
x =X ,
3
x = X 3,
where

(1.7)
Figure 1.3.

x =

X i,

this continuum

a(t)

is a given function. In

Simple

shear

of

beam

has become a parallelepiped, all

cross-sections of which are planes orthogonal to


the

Ox3

axis and are the same parallelograms.

This motion law is called simple shear; the tangent


of the shear angle

a.

is equal to

Example 1.3. Consider a continuum

B,

which in

is a

rectangular parallelepiped (a beam) shown in Figure 1.4;

under the transformation from

to

this parallelepiped

changes its dimensions without a change in its angles (as


in Example 1.1) and rotates by an angle

Ox1 x2

around the point

(t)

in the plane

(Figure 1.4). The motion law

for the continuum is called the rotation of a beam with

beam with extension

extension. In this case equations (1.1) have the form

xi = F0 ij xj ,
F0 ij is
matrix U0 :

where the matrix


the stretch

F0 ij

Figure 1.4. Rotation of a

xj = X j ,

(1.8)

the product of two matrices, the rotation matrix

= O0

k
k U0 j ,

O0

cos sin
= sin cos
0

0
0

O0

and

Chapter 1. Kinematics of Continua

U0
and

k1

k2

k (t) = h (t)/h0

F0 ij

k3

k1 cos k2 sin
= k1 sin k2 cos
0

k3

are the proportion functions characterizing the ratio of

lengths of the beam edges in

and

(as in Example 1.1).

1.1.2. Material and Spatial Descriptions. In continuum mechanics, physical


processes occurring in bodies are characterized by a certain set of varying scalar
fields

= (M, t),

a = a(M, t),

vector fields

and tensor fields of the

nth

order

n (M, t). We will consider tensors and tensor fields in detail in paragraph 1.1.4
(see also [12]).
Since in the Cartesian coordinate system
to both Lagrangian coordinates

Xi

O
ei

a material point

and Eulerian coordinates

xi ,

M corresponds
varying scalar

and vector fields can be written as follows:

e j , t) = (x
e , t),
(X i , t) = (X i (xj , t), t) = (x
i
j
e(x , t) = a
e(x, t),
a(X , t) = a

(1.9)

With the help of the motion law (1.3) (or (1.4)), we can pass from functions
of Lagrangian coordinates to functions of Eulerian coordinates in formulae (1.9).
In continuum mechanics the tilde
For a fixed time

is usually omitted (we will do this below).

in (1.9), we obtain stationary scalar and vector fields.

M is fixed, and
6 t 6 t0 , then we get an ordinary scalar
function a = a(M, t) depending on time.

If in (1.9) a material point

time

function

changes within the interval

= (M, t)

and vector

According to relationships (1.9), there are two ways to describe different


physical processes in continua.
In the material (Lagrangian) description of a continuum, all tensor fields
describing physical processes are considered as functions of

Xi

and

t.

In the spatial (Eulerian) description, all tensor fields describing physical


processes are functions of

xi

and

t.

Both the descriptions are equivalent. It should be noted that for solids we
more often use the material description, where it is convenient to fix coordinates

Xi

of a material point

and to observe its motion at different times

t.

For

gaseous and fluid continua, Eulerian description is more convenient; when an

xi and monitor the material


i
points M passing through this point x at different times t.

1.1.3. Local Bases in K and K. Using the motion law (1.4) and relationship
i
(1.1), at every material point M with coordinates X in the actual and reference
observer fixes a geometric point with coordinates

configurations we can introduce its local basis vectors:

rk =
where

xi
x
i = Qi k e
i ,
=
e
X k
X k

rk =

x
xi
i

i ,
=
e
=
Q
i
ke
X k
X k

Qi k = xi /X k ,

Qi k = xi /X k ,

P ik = X i /xk ,

P i k = X i / xk

(1.10)

(1.11)

1.1. Material and Spatial Descriptions

are Jacobian matrices and inverse Jacobian matrices.


Here and below all values referred to the configuration

will be denoted by

. As follows from the definition (1.11), local bases vectors r and


k

rk are directed tangentially to corresponding coordinate lines X k (Figure 1.5)

superscript

Figure 1.5. Local basis vectors in reference and actual configurations

In

and

introduce metric matrices

kl
g kl , g as follows:

gkl , g kl

and inverse metric matrices

xi xj
ij ,
X k X l

gkl = rk rl = Qi k Qj l ij =

xi xj
g kl = rk rl =
ij ,
X k X l

k
g kl glm = m
,

(1.12)

k
g kl g lm = m
,

and also vectors of reciprocal local bases

ri = g im rm ,

ri = g im rm ,

(1.13)

which satisfy the reciprocity relations

ri rj = i j ,

r i r j = i j ,

and also the following relations:

rn rm

= g nmk rk ,

rn rm

= g nmk rk ,

(1.14a)

(1.14b)

With the help of the mixed multiplication, i.e. sequentially applying scalar
and vector products to three different local bases vectors, we can determine the
volumes

|V |

and

|V |

constructed by these vectors:

|V | = r1 (r2 r3 ) = g = det (gij ) = xk /X i ,

|V | = r1 (r2 r3 ) = g = |xk /X i |.

(1.15)

Chapter 1. Kinematics of Continua

It should be noted that although local bases


in different configurations

and

(if one consider the same point

a(M)

and

ri

have been introduced

K, they correspond to the


M); therefore each of the

as a rigid whole into the same point in


any vector field

ri

or in

K.

same coordinates

bases can be carried

Due to this, we can resolve

i i
for each of the bases ri , r , r and ri :

a = ai ri = ai ri = ai ri = ai ri .
If curvilinear coordinates

onal as well: (ri

rj = ij ),

Xi

orthonormal basis

(1.16)

are orthogonal, then the vectors

ri

are orthog-

g ij and g ij are diagonal; hence we


q

H = g ( = 1, 2, 3) and the physical

and matrices

can introduce Lam


e's coefficients


r = r /H = r H .

Components of a vector

(1.17)

with respect to this basis are called physical:

b b
a = ai ri .
The actual basis

Xi

ri

(1.18)

is in general not orthogonal even if the basis

ri

orthogonal; therefore we cannot introduce the corresponding physical basis in


One can introduce a physical basis in

not with the help of

ri ,

is

K.

but with the

help of another special basis (see paragraph 1.1.7).

1.1.4. Tensors and Tensor Fields in Continuum Mechanics. For different

i at every point M, and with the help of formulae


ri , ri , ri , ri or e
given in the work [12] we can introduce different dyadic (tensor) bases: ri rj ,

ri rj , ri rj , ri rj , ri rj etc., which are equivalence classes of vector sets


consisting of 2 3 = 6 vectors (for example, r1 rj = [r1 rj r2 0r3 0], where [ ] is
local bases

the notation of an equivalence class), and


field of second-order tensor

is the sign of tensor product. A

T(M)

can be represented as a linear combination of

dyadic basis elements:

i e
j = T ij ri rj .
T = T ij ri rj = T ij ri rj = Tij e
During the passage from one basis to another, tensor components

(1.19)
T ij

are

transformed by the tensor law:

Tij = P ik P jl T kl = P i k P j l T kl .
Metric matrices
tensor

g im , gim , g im

and

g im

(1.20)

are components of the unit (metric)

with respect to different bases:

E = gim ri rm = g im ri rm = g im ri rm = g im ri rm .
For
the

(1.21)

tensor T, in continuum mechanics one often uses


TT = T ij rj ri and the inverse tensor T1 , where

second-order

transpose

T1 T = E.

tensor

The inverse tensor exists only for a nonsingular tensor (when

1.1. Material and Spatial Descriptions

det T 6=

0). The determinant of a tensor is defined by the determinant of its

det T = det T ij .

mixed components matrix:

Besides second-order tensors, in continuum mechanics one sometimes uses


tensors of higher orders [12]. To introduce the tensors, we define polyadic bases
by induction:
consisting of

ri . . . rin ; the
n 3 = 3n vectors.

bases are equivalence classes of vector sets

n-th

A field of

n (M) can be

order tensor

represented by a linear combination of polyadic basis elements:

n
where

...in

= i

and

...in

ri . . . rin = i

...in

...in

ri . . . rin ,

are components of the

n-th

order tensor with respect

are the first, second and

to the corresponding polyadic basis.


For fourth-order tensors, analogs of the tensor
third unit tensors defined as follows:

I = ei ei ek ek = E E,
II = ei ek ei ek , III = ei ek ek ei ,

(1.22)

and also the symmetric fourth-order unit tensor


1

= (II + III ).

(1.22a)

ijkl = ( ik jl + il jk ).

= ijkl ei ej ek el ,

We can transpose fourth-order tensors as follows:


4

where

(m

m2 m3 m4 )

= i

i i i

1 2 3 4

(m1 m2 m3 m4 ) is some
ri ri ri .

i1 i2 i3 i4 ri4

rim rim rim rim


3

substitution,

for

example,

(4321) =

1.1.5. Covariant Derivatives in


operators in configurations

and

= rk

and

K,

K.

Introduce the following nabla-

respectively:

,
X k

= rk

.
X k

(1.23)

Applying the nabla-operators to a vector field, we get the gradients of a


vector in

and

K:
a = rk

a = rk

a
= k ai rk ri ,
X k

(1.24)

a
= k ai rk ri = k ai rk ri = k ai rk ri ,
k
X

where we have denoted the following covariant derivatives in different tensor

bases in configurations

k ai =

and

K:


ai
m
ik am ,
k
X

k ai =

ai
+ ikm am
k
X

(1.25)

Chapter 1. Kinematics of Continua

ai
m
ik am ,
X k

k ai =
Here

m
ij

k ai =

ai
+ ikm am .
X k

m
ij

and

are the Christoffel symbols in configurations

K.

and

For the

Christoffel symbols the following relations (see [12]) hold:

gki
g
ijk ,
j
2
X
X
X !

g ij

g ki
1 km
kj
m
=
g
+

.
ij
2
X i
X j
X k
1

km
m
ij = g

Contravariant derivatives in

ai = g

(1.26)

and

km

kj
i

are introduced as follows:

k ai = g km m ai .

m ai ,

(1.27)

The covariant derivatives (1.25) are components of the second-order tensors

a and a,

therefore during the passage from the local basis

ri

to another

one they are transformed by the tensor law (1.20).

The nabla-operators

n (X i ) of

n-th

and

in

and


n
= k i
k
X
n
n
k
=r
= k i
k
X

k i

where

can be applied to a tensor field

order:

= rk

...in

and

k i

...in

...in k

r ri . . . rin ,

(1.28)

...in k

r ri . . . rin ,
1

are the covariant derivatives in

and

K,

respectively:

k i

...in

X k

...in

n
X

s
imk
i

...is =m...in

(1.29)

s=1

In the same way we can define operations of scalar product of the nabla-operator

in

(the divergence of a tensor):

= rk


n
= k ki
k
X

...in

ri . . . rin ,

(1.30)

and vector product of the nabla-operator in

= rk


1
=q
k

X
g

M0

and

M0

dx

ri . . . rin .
2

(1.31)

Consider how a local neighborhood of a

is transformed during the passage from configuration

arbitrary elementary radius-vector


points

(the curl of a tensor):

ijk i ji2 ...in rk

1.1.6. The Deformation Gradient.


point

connecting in

(Figure 1.6). In configuration

K,

are connected by the elementary radius-vector

to

K.

Take an

two infinitesimally close

these material points

dx.

and

1.1. Material and Spatial Descriptions

Figure 1.6. Transformation of an elementary radius-vector during the passage from the reference
configuration to the actual one

The vectors

dx

and

dx

can always be resolved for local bases:

x
dx(X ) =
dX k = rk dX k ,
k
X
On multiplying the first equation by

dX m

of

dx = rk r dx.

(1.33)

rm

and the second  by

rm dx = rm rk dX k = dX m ,
Substitution

x
dx(X ) =
dX k = rk dX k .
k
X

rm ,

we get

rm dx = rm rk dX k = dX m .
into

the

first

equation

of

(1.32)

(1.32)

(1.33)
yields

Changing the order of the tensor and scalar products (that is

permissible by the tensor analysis rules), we get the relation between

dx:

dx = F dx.

dx

and

(1.34)

Here we have denoted the linear transformation tensor

F = rk rk ,

(1.35)

called the deformation gradient. As follows from (1.34),


the deformation gradient connects elementary radiusvectors

dx

and

configurations

dx

of the same material point

and

in
Figure 1.7. Geometric rep-

K.

Definition (1.19) allows us to give a geometric


representation of the deformation gradient: if
considered as the left vectors and

ri

tion gradient

are

 as the right vectors, then, by formulae

of paragraph 1.1.4 (see [12]), the tensor

ri

resentation of the deforma-

takes the form

F = ri ri = [r1 r1 r2 r2 r3 r3 ].
According to the geometric definition of a tensor (see paragraph 1.1.4), the
tensor

ri , ri

can be represented as equivalence class of the ordered set of six vectors

(Figure 1.7).

10

Chapter 1. Kinematics of Continua

Besides

F,

in continuum mechanics one often uses the transpose tensor

1 and the inverse to the transpose tensor


the inverse tensor F

FT = rk rk = rk
1T

x
= x,
k
X

F1T :

F1 = rk rk ,

FT ,

(1.35a)

x
= x.
= (rk r ) = r rk = r
k
X

k T

It follows from (1.35) that

F ri = rk rk ri = rk ik = ri .

(1.36)

i.e. the deformation gradient transforms local bases vectors of the same material
point

from

Theorem 1.1.

to

K.

The transpose deformation gradient

an arbitrary vector

in

and

connects gradients of

K:

a = FT a,
H

FT

a = F1T a.

(1.37)

To derive formulae (1.37), we apply the definitions (1.24) and (1.35):

a = ri

a
a
a
= rj ji
= rj rj ri
= F1T a. N (1.38)
i
i
i
X
X
X

1.1.7. Curvilinear Spatial Coordinates.


basis

O
ei

Notice that the choice of Cartesian

as a fixed (immovable) system in the spatial (Eulerian) description

of the continuum motion is not a necessary condition. For some problems of

0i with
O0 e

0
0
0i
the origin at a moving point O (x0 = OO ) and a moving orthonormal basis e
i by the orthogonal tensor Q:
(Figure 1.8), which is connected to e

continuum mechanics it is convenient to consider a moving system

0i = Q e
i .
e

Figure 1.8. Moving bases

0i
e

and

e
ri

(1.39)

and curvilinear spatial coordinates

In this case, instead of Cartesian coordinates

xi of
0i :
e

one consider its Cartesian coordinates x


ei in basis

e = x x0 = x
0i .
x
ei e

ei
X

in moving system

a point

0i
O0 e

in the basis

i
e

(1.40)

11

1.1. Material and Spatial Descriptions

Let

Qi j

be components of the tensor

with respect to the basis

i :
e

i e
j ,
Q = Qi j e

(1.41)

then relation (1.39) takes the form

0i = Qj i e
j ,
e
x
ei

and coordinates

and

xi

(1.42)

are connected as follows:

e = x x0 = (xi xi0 )
0i = x
j ,
x
ei = x
ei e
ei Qj i e
xi xi0 = Qi j x
ej ,
i

x /e
x =Q

x
ei , we
0
point O :

nates

with the origin at

(1.43)
i

e
x /x =

j,

Instead of Cartesian coordinates

ek
X

P ji

can consider special curvilinear coordi-

e k ),
x
ei = x
ei (X

(1.44)

i
which, due to (1.27), are connected to x by the relations

e k ) xi (X
e k , t)
xi = xi0 (t) + Qi j (t)e
xj (X

ej = X
e j (xi , t).
X

or

(1.45)

xi0 (t) and Qi j (t)


0
0
i ), which are assumed to be known in
(i.e. only by the motion of system O e
The dependence on

in the relations is defined by functions

continuum mechanics.

ei
X

Coordinates

are no longer Lagrangian (material): at different times they

correspond to different material points. However, it is often convenient to choose


coordinates

ei
X

Xi

coincident with

in the reference configuration

K.

In this case

we have the relations

e j , 0).
xi (X i ) = xi (X j , 0) = xi (X

(1.46)

With the help of transformation (1.45) we can use the spatial description in
coordinates

ei
X

as well when consider the functions

e i , t);
e(X
a = a(xi , t) = a
therefore coordinates

ei
X

(1.47)

are called curvilinear spatial coordinates.

Introduce local vectors

e
ri =
In particular, the basis

0i
e

is independent of

coincides with

ej = X
e j (xi )
X

0i = e
i , x
e = x,
e
independent of t; the basis

are

as well, and from (1.46) and (1.48) it follows that the basis

ri :

e
ri =
When the basis

(1.48)

may be fixed (Figure 1.9), then

and curvilinear spatial coordinates

e
ri

x
xj
.
=
e
ei
ei j
X
X

0i
e

xj
xj

= ri .
e
=
e
j
i j
i
e
X
X

is moving, bases

e
ri

and

ri

(1.49)

are no longer coincident.

12

Chapter 1. Kinematics of Continua

Figure 1.9. Curvilinear spatial coordinates

The vectors

e
ri

e i and Lagrangian
X
0i = e
i
e

coordinates

Xi

for the fixed basis

are directed tangentially to the coordinate lines

simultaneously with

ri

at every point

A change of vectors

e
ri

at any time

ei
X

and defined

t > 0.

in time is defined only by the motion of basis

0i ,
e

because from (1.42), (1.43) and (1.48) it follows that

0i = Qj i
e

ek
ek
X
xj X
e
e
r
=
rk =
k
xj
e
xi xj

X
ek
e
rk ,
i

e k /e
Peki X
xi is independent of t according
vectors ri and e
ri are connected as follows:

and the matrix


The bases

ri =

to (1.44).

ek
ek
x
x X
X
e
rk .
=
=
i
i
e k X
X
X i
X

Just as in paragraph 1.1.2, we define the metric matrix


metric matrix

(1.50)

e
x

geij :

geij = e
ri e
rj ,

(1.51)
geij

geij gejk = ki ,

and the inverse

(1.52)

and the reciprocal basis vectors

e
ri = geike
rk =

ei j
e i 0k
X
X
=
.
e
e
j
x
e
xk

(1.53)

According to formulae (1.51) and (1.52), we find the relation between


matrices

gij

and

gekl :
gij = ri rj =

The inverse matrix

g ij

e k X
el
e k X
el
X
X
e
r e
rl =
gekl .
i
j k
i
X X
X X j

(1.54)

is found from (1.54) by the rule of matrix product

inversion (see Exercise 1.1.13):

g ij =

X i X j kl
ge .
e k X
el
X

(1.55)

13

1.1. Material and Spatial Descriptions

From (1.51), (1.53) and (1.55) we can find the relation between vectors of
reciprocal bases

ri

e
ri :

and

ri = g ij rj =
Let there be a tensor
basis

e
ri :

= i

em
X i X j kl X
X i k
e
e
rm =
ge
r .
j
k
l
e X
e
ek
X
X
X

(1.56)

n , then it can be resolved for the basis

...in

ei
ri . . . rin =

...in

ri

and for the

e
rin .
ri . . . e

(1.57)

On substituting (1.51) into (1.57), we derive transformation formulae for


tensor components during the passage from coordinates

ei

...in

Introduce the nabla-operator

= j

...jn

ei
X
X j

...

Xi

to

e i:
X

e in
X
.
X jn

(1.58)

of covariant differentiation in coordinates

e =e

ri

(1.59)

ei
X

and contravariant derivatives of components

e i:
X

e
ai
e ke
eme

ai =

ik am ,
k
e
X

The Christoffel symbols

em

ij

e
ai

of a vector

ri
a=e
aie

in coordinates

e
ai
m
e ke
ei e

ai =
+
km a .
k

(1.60)

e
X

in coordinates

ei
X

are connected to

geij

relations which are similar to (1.26).

Theorem 1.2. The results of covariant differentiation


X i (in the configuration K) are coincident:

e n ,
n =
H

e i:
X

e n ,
n =

in coordinates

e n .
n =

by the

ei
X

and

(1.61)

Prove the first formula in (1.61). Due to (1.23), we have

n = ri

el
n X
n
X i k
n
l k
e n . (1.62)
e
=
r

=
k
i
i
k
l
l
e
e
e
X
X
X X
X

The remaining two formulae in (1.61) can be derived in the same way (see
Exercise 1.1.8).

Going to components of a tensor

n with respect to bases

ri

and

e
ri ,

from

(1.58) we get the relation between the covariant derivatives:

i j

Determine the tensor

...jn

e i
ej
=

...jn

transforming coordinates

(1.63)
ei
X

into

Xj:

e ij e
ij e
i e
j .
H = rj e
rj = H
ri e
rj = H

(1.64)

Then we get the relations (see Exercises 1.1.10 and 1.1.11):

ej e
ri = H e
ri = H
i rj ,

e ij )1e
ri = H1T e
ri = (H
rj .

(1.65)

14

Chapter 1. Kinematics of Continua

ei
X

e
ri and e
ri
ij
are orthogonal as well, and the matrices g
eij and ge are diagonal; and we can
The coordinates

are often chosen orthogonal, then the bases

introduce the physical (orthonormal) basis:

where

e =
H

b
e,
e
r = e
r /H

ge

are Lam
e's coefficients, which are in general not coincident

with the coefficients

b
e
r

(1.66)

H =

g .

are called physical:

Tensor components with respect to the basis

b
e
e
T = Teij b
ri b
rj .

(1.67)

Relations between physical and covariant components of a tensor are determined


by the known formulae (see [12]).

Exercises for 1.1.


Exercise 1.1.1. With the help of formulae (1.10), (1.12), (1.13) and (1.17) show that if
the motion law of a continuum describes extension of a beam (1.6) (see Example 1.1),
then the local basis vectors

ri

and the metric matrices have the forms

r i = ei ,

ri = ei ,

r = (1/k )
e ,

,
r = k e

g ij = ij ,

2
k1 0 0
(gij ) = 0 k22 0 ,
0
0
k32
i.e.

= 1, 2, 3,

ij

ij

= ,
2
k1
0
(g ij ) = 0 k22
0

g = k2 , g = k2 , H =

0
0

k3

g = k , b
r = e .

Exercise 1.1.2.

Show that if the motion law of a continuum describes a simple shear


(see Example 1.2), then the local basis vectors and the metric matrices have the forms

r i = ei ,

r i = ei ,

1 ,
r1 = e

g ij = ij ,

g ij = ij ,

2 ,
r2 = a
e1 + e

3 ,
r3 = e

a
,
3 ,
r =e
e ,
r =e
r3 = e

2
1
a
0
1+a
a
ij
2
gij = a 1 + a 0 ,
g =
a
1
1

0
0

!
.

Exercise 1.1.3. Show that if the motion law describes rotation of a beam with extension
(see Example 1.3), then with introducing the rotation tensor
U0 :

i e
j ,
O0 = O0i j e

U0 =

3
X

e
,
k e

=1
we can rewrite the beam motion law in the tensor form

x = F0 x,

O0

F0 = O0 U0 .

and the stretch tensor

15

1.1. Material and Spatial Descriptions

Show that the local basis vectors and metric matrices for this problem have the forms

k ,
ri = F0 ki e
gij = F0 ki F0 lj kl

i ,
ri = e

2
k1 cos2 + k22 sin2 (k12 k22 ) cos sin
= (k12 k22 ) cos sin k12 sin2 + k22 cos2
0

g = k1 k2 k3 ,

g ij

k2 sin + k1 cos2 (k12 k22 ) cos sin

= (k12 k22 ) cos sin k22 cos2 + k12 sin2


2

0
0

2
3

k3

Exercise 1.1.4.

Using the property (1.14) of reciprocal basis vectors, show that the
following relations hold:

ri =
Exercise 1.1.5.

Show that

X i
k

k =
ri = P ik e

k ,
e

F, FT , F1

and

F1T

X i k
.
e
xk

in the Cartesian coordinate system

take the forms

F=

xm
i

m e
i ,
e

FT =

X k
i

i
e

xm
xm i
= i e
e
m ,
e
k m
X
x

F1 =

xm
X k i
xm

=
m ei ,
e
e
e
i
k m
x
xi
X

F1T =

X k i
xm
xm i

=
em .
e
e
e
i
k m
x
xi
X

Exercise 1.1.6. Substituting (1.54), (1.52) and (1.55) into (1.12), derive formula (1.55).
Exercise 1.1.7. Prove that

ri = F1T ri .

Exercise 1.1.8. Derive the third formula of (1.61).


Exercise 1.1.9. Prove that for any scalar function
connected by the relationship

(X i )

its gradients in

and

are

= F1T .
Exercise 1.1.10. Show that formulae (1.65) follow from (1.64).
Exercise 1.1.11.

Using (1.47), show that in formulae (1.64) the tensor


i and e
following components with respect to bases e
ri :

i
ek
ij = x X ,
H
j
k

ij )1 =
(H

k
ei
e ij = x X ,
H
j
k

xi X k
j ,
e k x
X

X x

e ij )1 =
(H

has the

xk X i
k .
e j x
X

Exercise 1.1.12.
gradient

Introducing the notation

with respect to basis

(1.36) yields

F ij

ij

for components of the deformation

ri : F = F ri rj = F i j ri rj ,

rj = F i j ri .

show that formula

16

Chapter 1. Kinematics of Continua

Exercise 1.1.13. Show that the Levi-Civita symbols are connected by the relations

ijk ijk = 6,

where

T jk

ijk ilm = jl km kl jm ,

ijk ijl = 2il ,

g ijk = (1/ g ) mnl gmi gnj glk .

ijk T jk = 0

are components of an arbitrary symmetric tensor:

T jk = T kj .

Exercise 1.1.14.

Using relations (1.14a), show that the local bases vectors are
connected by the relations

r r =

g r ,

g r ,

r r =

6= 6= 6= .

Exercise 1.1.15. Show that the unit fourth-order tensors

I , II

and

III

defined by

formulae (1.22) have the following properties:

I T = I1 (T)E,

I1 (T) = T E,
1

T = (T + TT ),

III T = T,
and

II T = TT ,

I 4 = E E 4 ,

II 4 = (2134) ,
1

4 = ((2134) + ),

III 4 = 6 ,

for arbitrary second-order tensor T and fourth-order tensor . As follows from these
formulae, the tensor III is the `true' unit fourth-order tensor.
4

Exercise 1.1.16. Show that components of the symmetric unit fourth-order tensor
with respect to a tetradic basis have the form

= (ei el ei el + ei el el ei ) = ijkl ei ej ek el .
2

ijkl = ( ik jl + il jk ).
2

Exercise 1.1.17. Show that for any second-order tensor

and for any vector

the

following formula of covariant differentiation hold:

(T a) = T ( a)T + a T.

1.2. Deformation Tensors and Measures


1.2.1. Deformation Tensors.

Besides

F,

important characteristics of the

motion of a continuum are deformation tensors, which are introduced as follows:

C = (gij g ij )ri rj = ij ri rj ,
2

A = (gij g ij )ri rj = ij ri rj ,
2

= (g ij g ij )ri rj = ij ri rj ,
2

(2.1)

17

1.2. Deformation Tensors and Measures

J = (g ij g ij )ri rj = ij ri rj .
2

Here

is called the right CauchyGreen deformation tensor,

Almansi deformation tensor,

 the left

 the right Almansi deformation tensor, and

 the left CauchyGreen tensor.


As follows from the definition of the tensors, covariant components of

and

are coincident, but they are defined with respect to different tensor bases.

Components

ij

are called covariant components of the deformation tensor.

Contravariant components of the tensors


called contravariant components

ij

and

are also coincident and

of the deformation tensor, but they are

defined with respect to different tensor bases of the tensors

and

J.

Notice that the deformation tensor components

ij = (g ij g ij ),

ij = (gij g ij ),
2

(2.2)

have been defined independently of each other, therefore the formal rearrangement of indices is not permissible for these components, i.e.

kl = ij g ik g jl 6= kl ,

kl = ij gik gjl 6= kl .

(2.3)

Thus, when there is a need to obtain contravariant components from


covariant components from

ij ,

one should use the notation

kl

and

kl .

ij

and

We will

also use the notation

kl = ij g ik g jl ,
Theorem 1.3.

The deformation tensors

deformation gradient

kl = ij g ik g jl .
C, A,

and

(2.4)
are connected to the

as follows:

C = (FT F E),

A = (E F1T F1 ),

= (E F1 F1T ),
2

J = (F FT E).
2

(2.5)

H Let us derive a relation between C and F. Having used the definitions of gij ,

g ij and F, we get
1
1

1
C=
(ri rj )ri rj E =
ri ri rj rj E = (FT F E).
2
2
2
(2.6)
The remaining relations of (2.5) can be proved in the same way:

A = (E ri ri rj rj ) = (E F1T F1 ),
2

= (E ri ri rj rj ) = (E F1 F1T ),
2

J = (ri ri rj rj E) = (F FT E). N
2

(2.6a)

18

Chapter 1. Kinematics of Continua

1.2.2. Deformation Measures.

Besides the deformation tensors, we define

deformation measures: the right CauchyGreen measure


measure

g:

G and the left Almansi

G = gij ri rj = FT F = E + 2C,

g = g ij ri rj = F1T F1 = E 2A,
and also the left CauchyGreen measure

G1 :

g1

(2.7)

and the right Almansi measure

g1 = g ij ri rj = F FT = E + 2J,

G1 = g ij ri rj = F1 F1T = E 2.
1.2.3. Displacement Vector. Introduce a

displacement vector

(2.8)
u

of a point

from the reference configuration to the actual one as follows (Figure 1.10):

u = x x.

Figure 1.10. The displacement vector of a point

(2.9)

from the reference configuration to the actual

one

Theorem 1.4.

The

deformation

tensors

connected to the displacement vector

F = E + ( u)T ,

FT = E + u,
and also

C=

A=

and

the

deformation

F1T = E u,

(2.10)

T
T
u+u +uu
,

are

F1 = E ( u)T ,

u + uT u uT ,

u + ( u)T uT u ,
2

1
T
T
u+u +u u .
J=

gradient

by the relations

(2.11)

19

1.2. Deformation Tensors and Measures

The definition (2.9) of the displacement vector and the properties (1.35) of the

deformation gradient yield

x
+u=
i
X

FT = x = (x + u) = ri

= ri ri + u = E + u.

(2.12)

C takes the form

1
T
C=
(E + u) (E + u ) E =
2

1
T
T
=
u+u +uu
.

(2.13)

Then the tensor

In a similar way, we can prove the remaining relations of the theorem.

1.2.4. Relations between Components of Deformation Tensors and Displacement Vector. The displacement vector u can be resolved for both bases

ri and ri :

u = ui ri = ui ri .
(2.14)

Xi

The derivative with respect to


well:

u
=
X i

can be determined in both the bases as


i uk rk

= i uk rk .

(2.15)

Then the displacement vector gradients take the forms

u
= i uk ri rk = i uk ri rk ,
i
X
u
u = ri
= i uk ri rk = i uk ri rk .
X i

u = ri

(2.16)
(2.17)

Substitution of these expressions into (2.10) gives

F = (ik + i uk )rk ri = F ki rk ri .

(2.18)

Here we have introduced components of the deformation gradient in the reference


configuration:

F ki = ik + i uk .

(2.19)

T
The transpose gradient F has the components

FT = F ki ri rk = F i k rk ri ,

F i k = ik + k ui ,

where

(2.20)
(2.21)

(F i k )T = F ki .

In a similar way, one can find the expression for the inverse gradient

F1 = (ik i uk )rk ri = (F 1 )ki rk ri

(2.22)

20

Chapter 1. Kinematics of Continua

and for the inversetranspose gradient

F1T = (F 1 )ki ri rk = (F 1 )i k rk ri ,

(2.23)

where their components with respect to the actual configuration are expressed
as follows:

(F 1 )i k = ik k ui ,

(2.24)

(F 1 )ki = ik i uk .

(2.25)

Thus, we have proved the following theorem.

Theorem 1.5.

Components of the deformation gradients

in local bases of configurations


displacement vector

and

F, FT , F1

and

F1T

are connected to components of the

by relations (2.19), (2.20), (2.24) and (2.25).

On substituting formulae (2.16) and (2.17) into (2.11) for

and

and

comparing them with (2.1), we get


1

ij = (i uj + j ui + i uk j uk ),
2

ij = (i uj + j ui i uk j uk ),

(2.26)

 the expressions for covariant components of the deformation tensor in terms

of components of the displacement vector with respect to

and

K.

In a similar way, substituting (2.16) and (2.17) into (2.11) for


obtain

1 i j


j i

k i

and

J,

we

ij = ( u + u + u k uj ),
2

ij

= (i uj + j ui k ui k uj )

(2.27)

 the relations between contravariant components of the deformation tensor and

components of the displacement vector in

and

K.

Then with using relations (2.2), (2.26) and (2.27), we can find the connection
between the metric matrices and displacement components:

gij = g ij + i uj + j ui + i uk j uk = g ij + i uj + j ui i uk j uk , (2.28)

g ij = g ij + i uj + j ui + k ui k uj = g ij + i uj + j ui k ui k uj . (2.29)
Thus, we have proved the following theorem.

Theorem 1.6. Components of the deformation


ij are connected to components of
trices gij , g

tensor

ij , ij

and metric ma-

the displacement vector

by

relations (2.26)(2.29).

1.2.5. Physical Meaning of Components of the Deformation Tensor.

Let

us clarify now a physical meaning of components of the deformation tensor:


1

ij = (gij g ij ) = (ri rj ri rj ).
2

(2.30)

21

1.2. Deformation Tensors and Measures

By the definition of the scalar product (see [12]), we have


1

=
where

and

K,

|r ||r | cos |r ||r | cos ,

and

are the angles between basis vectors

r , r

(2.31)
and

r , r

in

respectively.

Consider elementary radius-vectors


and introduce their lengths

ds

and

ds,

dx

and

in configurations

and

K,

respectively:

ds2 = dx dx,

dx

ds2 = dx dx.

(2.32)

dx is arbitrary, we can choose it to be oriented along one of the basis

vectors r . Then dx will be directed along the corresponding vector r as well,

because under this transformation r becomes r for the same material point M
k
with Lagrangian coordinates X . In this case we have
Since

x
|dx| = ds = dX = |r | dX ,
X

|dx| = ds = dX = |r | dX .

(2.33)

Hence

where

ds /ds = |r |/|r | = + 1,

(2.34)

is called the relative elongation. Formula (2.34) yields

|r | = |r |(1 + ).

(2.35)

On substituting this expression into (2.31), we get

= |r ||r | (1 + )(1 + ) cos cos .


1

(2.36)

Consider the case when

= ,

then

= = 0

and

1
= |r |2 (1 + )2 1 = (1 + )2 1 .
2

Let coordinates

= ;

be coincident with Cartesian coordinates

xi ,

and for infinitesimal values of the relative elongation, when

obtain

i.e.

Xi

(2.37)

then

1,

,
is coincident with the relative elongation.

In general,

g =

is a nonlinear function of corresponding elongations.

we

(2.38)

22

Chapter 1. Kinematics of Continua

Consider

6=

and assume that

X i = xi ,

= /2;

then

and from (2.36)

we get

= |r ||r |(1 + )(1 + ) cos =


2
q
q

1
1
=
g g (1 + )(1 + ) sin = (1 + )(1 + ) sin ,
2

(2.39)

= = (/2) is the change of the angle between basis


vectors r and r . For small relative elongations, when 1, and small angles
1, from (2.39) we get
where

/2,
i.e.

(2.40)

is a half of the misalignment angle of the basis vectors.

1.2.6. Transformation of an Oriented Surface Element.


ration

and X .

consider a smooth surface

In actual configu-

which contains two coordinate lines

Then we can introduce the normal

as follows:

to the surface

n = p r r .

(2.41)

ge

Here

ge = det (e
g ),

and

matrix of the surface (,

ge is the
= 1, 2):

two-dimensional

ge = r r
Figure 1.11. Introduction of
oriented

surface

n d

element

In configuration

are directed along local basis vectors, i.e.

dx = r dX

p
d = ge dX dX

is called the area of the surface element

gij =

consider a surface element

constructed on elementary radius-vectors

dx .

(2.42)

(it is not to be confused with the metric matrix

= ri rj ).

dx ,

which

(Figure 1.11). The value

constructed on vectors

(2.43)
dx

and

Then formula (2.41) takes the form

n d = r dX r dX = dx dx ,
where

n d

(2.44)

is called the oriented surface element.

Show that the normal

defined by formula (2.41) is a unit vector. According

to the property (1.14b) of the vector product of basis vectors and the results of
Exercise 1.1.13, we can rewrite equation (2.44) in the form

n d = r r dX dX =

g r dX dX =

= (1/ g )ijk gi gj rk dX dX

(2.45)

23

1.2. Deformation Tensors and Measures

, ).

(there is no summation over

n d n d =

Thus,

1
g r ijk gi gj rk )(dX dX )2 =
g

= (k

ijk

gi gj )(dX dX )2 = (i j i j )gi gj (dX dX )2 =


2
= (g g g
)(dX dX )2 = ge(dX dX )2 = d2 .

Hence,

n n = 1.

The surface element

in

corresponds to the surface element

dx

which is constructed on elementary radiusvectors

(2.46)

and

(2.47)

is the unit normal to

Since

n d =

K,

dx :

n d = r dX r dX = r r dX dX .
Here

in

r = F1T r

d.

, we get

g F1T r dX dX =

Thus, we have proved the following theorem.

Theorem 1.7.

connected by the relation

g/g n F1 d =

n d

The oriented surface elements

n d =

g/g F1T r r dX dX =
q

= g/g F1T n d.

and

n d

and

in

g/g F1T n d.

(2.48)

are

(2.49)

With the help of the deformation measures we can derive formulae connecting
the normals
in

and

and

to the surface element containing the same material points

K.

Multiplying the equation (2.49) by itself and taking the formula

nn=

into account, we get

d2 = (n F1 F1T n) d2 = (n G1 n) d2 .
g

Thus,

d/d =

g/g (n G1 n)1/2 .

On the other hand, expressing

(2.50)

(2.51)

from (2.49) and then multiplying the

obtained relation by itself, we obtain

d 2 =
Thus, we find that

g
g
(n F FT n) d2 = (n g1 n) d2 ,
g
g

d/d =

g/g (n g1 n)1/2 .

(2.52)

(2.53)

24

Chapter 1. Kinematics of Continua

On introducing the notation

k = (n G1 n)1/2 = (n F1 F1T n)1/2 ,


k = (n g1 n)1/2 = (n F FT n)1/2 ,

(2.54)

relations (2.52) take the form

q
q

d/d = g/g k = g/g (1/k).

Thus, we get

(2.55)

k = 1/k.

(2.56)

Substitution of (2.52) and (2.53) into (2.49) gives the desired relations

kn = F1T n,

k n = FT n.

(2.57)

1.2.7. Representation of the Inverse Metric Matrix in terms of Components


of the Deformation Tensor. Components gij of the metric matrix are connected

ij

to components of the deformation tensor

by relation (2.2). In continuum

mechanics, one often needs to know the expression of the inverse metric matrix

g ij

in terms of

ij

(but not in terms of

ij ).

To derive this relation, we should

use the connection between components of a matrix and its inverse (see [12]):

g ij =
For

g ij ,

1
2g

imn jkl gmk gnl .

(2.58)

we have the similar formula

g ij =

imn jkl

2g

g mk g nl .

On substituting the relations (2.2) between

(2.59)

gmn , g mn

and

mn

into (2.59), we

get

g ij =

1
2g

imn jkl (g mk + 2mk )(g nl + 2nl ) =


=

1
2g

imn jkl (g mk g nl + 2g mk nl + 1g nl mk + 4mk nl ).

(2.60)

Removing the parentheses, modify four summands in (2.60) in the following way.
The first summand with taking formula (2.59) into account gives the matrix

g ij (g/g).

To transform the second and the third summands, we should use the

q
q
jkl


(1/ g ) = g tsp g jt g ks g lp .

formulae

(2.61)

(1/g)imn jkl g mk = imn tsp g jt g ks g lp g mk = imn tmp g jt g lp =

= (ti pn pi tn )g jt g lp = g ij g nl g jn g il .

(1/g)imn jkl g mk nl = (g ij g nl g jn g il )nl .

(2.62)

(2.63)

25

1.2. Deformation Tensors and Measures

Formula (2.61) follows from the relation

g ijk = (1/ g ) mnl gmi gnj glk


(see [12]), and relationship (2.62) has been obtained by using formula (2.61) and
the properties of the Levi-Civita symbols (see Exercise 1.1.13).
On substituting formula (2.63) into (2.60), we get

g ij =



g ij
2
g + 2(g ij g nl 2g il g jn )nl + imn jkl mk nl .
g
g

Finally, we should express the determinant


do this, we multiply relation (2.64) by
3g

gij

g = det (gij )

in terms of

(2.64)
ij .

To

and take formula (2.2) into account:




2
= g g ij + 2(g ij g nl g il g jn )nl + imn jpl mp nl (g ij + 2ij ).
g

(2.65)

Thus, we get
3g

= g(3 + 4g nl nl + (2/g)imn jpl g ij mp nl +

+ 2g ij ij + 4(g ij g nl g il g jn )nl ij + (4/g)imn jpl ij mp nl .

(2.66)

Modifying the third summand on the right-hand side by formula (2.63) and
introducing the notation

I1 = g nl nl ,

I2 = (g ij g nl g il g jn )ij nl ,
2

I3 = det (ij gik ),

(2.67)

from (2.66) we get the desired formula

g = g(1 + 2I1 + 4I2 + 8I3 ).

(2.68)

Here we have taken account of formula (2.58) for the matrix determinant
and also the relation

I3 = (1/g)det (ij ).

theorem.

Theorem 1.8.
nents

ij

The inverse metric matrix

of the deformation tensor and

ij

Thus, we have proved the following

g ij

is expressed in terms of compo-

by formulae (2.64) and (2.68).

Formulae (2.64) and (2.68) allow us to find the expression of contravariant


components

ij

of the deformation tensor in terms of

ij .

It follows from (2.2)

that

ij

ij

ij

ij g

= (1/2)(g g ) = g


g
g
1
(g ij g nl g il g jn )nl imn jkl mk nl .
2g
g
g

(2.69)

Substitution of formulae (2.26) or (2.27) into (2.64) and (2.69) gives the
expression for components

ui

or

ui .

ij

in terms of components of the displacement vector

26

Chapter 1. Kinematics of Continua

Exercises for 1.2


Exercise 1.2.1. Using the results of Exercise 1.1.1, show that the deformation gradient

and its inverse

F1

for the problem on a beam in tension (see Example 1.1) have the

forms

F=F =

3
X

e
,
k e

=F

1T

=1

3
X

=1

e
.
e

For this problem, the deformation tensors are determined by the formulae

C==

1
2

3
X

,
(k2 1)
e e

A==

=1

3
X

,
(1 k2 )
e e

=1

the deformation measures are determined as follows:

G = g 1 =

3
X

g = G1 =

e
,
k2 e

=1

3
X

e
,
k2 e

=1

and components of the deformation tensor take the forms


1

= (1 k2 ) .

= (k2 1) ,
2

Exercise 1.2.2.

Using the results of Exercise 1.1.2, show that for the problem on a

simple shear we have the following formulae for the deformation gradient:

i e
j = E + a
2 ,
F = F ij e
e1 e
2 ,
F1 = E a
e1 e

i.e.

ij

1 ,
FT = E + a
e2 e

1 ,
F1T = E a
e2 e

!
det F = 1,

for the deformation tensors:

C = (a/2)O3 + (a2 /2)


e22 ,

A = (a/2)O3 (a2 /2)


e22 ,

= (a/2)O3 (a2 /2)


e21 ,

J = (a/2)O3 + (a2 /2)


e21 ,

and for components of the deformation tensor:

(ij ) =

a/2

a/2
a2 /2

ij

( ) =

a/2

a/2
a2 /2

!
.

Here we have introduced the notation

1 e
2 + e
2 e
1 ,
O3 e

2 e
e
,
e

= 1, 2, 3.

Exercise 1.2.3. Using the formulae from Example 1.3 (see paragraph 1.1.1), show that
for the problem on rotation of a beam with extension, the deformation gradient has the
form

i e
j = cos
F = F0 ij e

2
X

=1

e
+ k3 e
3 e
3 + sin k2 (
1 e
1 e
2 ).
k e
e2 e

27

1.3. Polar Decomposition

Exercise 1.2.4. Using formulae (1.36), (2.18)(2.25) and the results of Exercise 1.1.7,
show that local basis vectors are connected to displacements by the relations

ri = rk (ik + k ui ),
Exercise 1.2.5.

Using formulae (2.37) and (2.39), show that the physical components

of the deformation tensor


and angles

ri = rk (ik i uk ).

b
= /

g g

are connected to relative elongations

by the relations

b
1
= (1 + )(1 + ) sin .

b
1
= ((1 + )2 1),
2

Exercise 1.2.6.

e
ri of curvilinear coordinate
u can be rewritten in the

Show that in the basis

expression of the tensor

in terms of

system

ei
X

the

form similar to

(2.22)(2.25):

rk e
ri ,
F1 = (Fe1 )ki e

u=u
eke
rk ,

e iu
(Fe1 )ki = ik
ek .

Exercise 1.2.7. Using formula (1.34), show that the following relationships hold:

|dx|2 = dx G dx,

|dx|2 = dx g dx.

1.3. Polar Decomposition


1.3.1. Theorem on Polar Decomposition. According to (1.36), the tensor F

can be considered as a tensor of the linear transformation from the basis ri to

the basis ri . Since the vectors ri and ri are linearly independent, the tensor F is
nonsingular. Then for this tensor the following theorem is valid.

Theorem 1.9 (on the polar decomposition).


tensor

F=OU
Here

Any nonsingular second-order

can be represented as the scalar product of two second-order tensors:

and

are

the

symmetric

or

F = V O.

and

positive-definite

(3.1)
tensors,

is

the

orthogonal tensor, and each of the decompositions (3.1) is unique.

Prove the existence of the decomposition (3.1) in the constructive way, i.e. we

should construct the tensors


of the tensor

U, V

O. To do this, consider
FT F and F FT . Both

and

with its transpose:

the contractions
the tensors are

symmetric, because

(FT F)T = FT (FT )T = FT F

and

(F FT )T = (FT )T FT = F FT , (3.2)

and positive-definite:

a (FT F) a = (a FT ) (F a) = (F a) (F a) = b b = |b|2 > 0


for any non-zero vector

(3.3)

a, where b = F a. Since any symmetric positive-definite


FT F

tensor has three real positive eigenvalues [12], eigenvalues of tensors

28

Chapter 1. Kinematics of Continua

F FT

and

can be denoted as

and

2 .

These tensors are diagonal in their

eigenbases, i.e. they have the following forms:

FT F =

3
X

2 p p ,

F FT =

=1
Here

3
X

2 p p .

(3.4)

=1

are the eigenvectors of the tensor

FT F

and

 of the tensor

F FT ,

which are real-valued and orthonormal:

p p = ,

p p = .

(3.5)

The right-hand sides of (3.4) are the squares of certain tensors

and

defined

as

U=

3
X

p p , > 0;

V=

=1
Here signs at

3
X

p p , > 0.

(3.6)

=1

are always chosen positive.

In this case, the following relations are valid:

FT F = U2 ,
The constructed tensors

and

F FT = V2 .

(3.7)

are symmetric due to formula (3.6) and

positive-definite, because for any nonzero vector

we have

3
3
X
X

aUa=
a p p a =
(a p )2 > 0,

=1

as

> 0.

In a similar way, we can prove that the tensor

Both the tensors

and

the theorem, the tensor

(3.8)

=1

is positive-definite.

are nonsingular, because, under the conditions of

is nonsingular. And from (3.7) we get

(det U)2 = det U2 = det (FT F) = (det F)2 6= 0.


Then there exist inverse tensors

U1

and

V 1 ,

(3.9)

with the help of which we can

construct two more new tensors

O = F U1 ,

O = V 1 F ,

(3.10)

which are orthogonal. Indeed,

OT O = (F U1 )T (F U1 ) = U1 FT F U1 = U1 U2 U1 = E.
(3.11)

According to [12], this means that the tensor


we can show that the tensor

is orthogonal. In a similar way,

is orthogonal as well.

Thus, we have really constructed the tensors

the product of which, due to (3.10), gives the tensor

F = O U = V O.

and

O,

and also

and

O,

F:
(3.12)

29

1.3. Polar Decomposition

Here

and

are symmetric, positive-definite tensors,

and

are orthogonal

tensors.
Show that each of the decompositions (3.12) is unique. By contradiction, let
there be one more resolution, for example

But then

hence,

e = U,
U

(3.13)

e 2 = U2 ,
FT F = U

(3.14)

because the decomposition of the tensor

is unique. Signs at
coincidence of

e U.
e
F=O

and

and

e
U

for its eigenbasis

are chosen positive by the condition. The

leads to the fact that

because

FT F

e
O

and

are coincident as well,

e =FU
e 1 = F U1 = O.
O

(3.15)

This has proved uniqueness of the decomposition (3.12). We can verify uniqueness of the decomposition

F=VO

in a similar way.

Finally, we must show that the orthogonal tensors

and

are coincident,

i.e. formula (3.1) follows from (3.12). To do this, we construct the tensor

F OT = O U OT .

(3.16)

Due to (3.12), this tensor satisfies the following relationship:

O U OT = V O OT .

(3.17)

The tensor

O OT

is orthogonal, because

(O OT )T (O OT ) = O OT O OT = O OT = E.

(3.18)

Then the relationship (3.17) can be considered as the polar decomposition of the

tensor

O U OT .

This tensor is symmetric, because

(O U OT )T = (OT )T (O U)T = O U OT .

(3.19)

Then the formal equality

O U OT = O U OT

(3.20)

is one more polar decomposition. However, as was shown above, the polar
decomposition is unique; hence, the following relationships must be satisfied:

V = O U OT
Thus, the orthogonal tensors
The tensors
tively, and

and

and

O OT = E.

and

(3.21)

are coincident:

O = O. N

are called the right and left stretch tensors, respec-

is the rotation tensor accompanying the deformation.

30

Chapter 1. Kinematics of Continua

The tensor

has nine independent components, the tensor

independent components, and each of the tensors

and

 three

 six independent

components.

Remark 1. Since the rotation tensor

is unique in the polar decomposition,

from formula (3.21) we get that the stretch tensors

V = O U OT ,
Theorem 1.10.

and

U = OT V O.

are connected by

(3.21a)

The CauchyGreen and Almansi deformation tensors can be

expressed in terms of the stretch tensors

and

as follows:

A = (E V2 ),

C = (U2 E),
2

= (E U2 ),

J = (V2 E).

O:

means of the tensor

(3.22)

To see this, let us substitute the polar decomposition (3.1) into (2.5), and then

we get the relationships (3.22).

1.3.2. Eigenvalues and Eigenbases.


Theorem 1.11. Eigenvalues of the tensors
coincident:

= ,
and eigenvectors

and

and

defined by (3.6) are

= 1, 2, 3,

(3.23)

are connected by the rotation tensor

O:

p = O p .
H

(3.23a)

To prove the theorem, we use the definition (3.6) and the first formula of

(3.21a):

V=

3
X

3
3
X
X

p p = O U O =
O p (O p ) =
p0 p0 ,

=1

=1

where

=1

p0 = O p .

According to the relationship, we have obtained two different eigenbases of the


tensor

and two sets of eigenvalues, that is impossible. Therefore,

p0 = O p = p
as was to be proved.

and

= ,

Due to (3.5), both the eigenbases are orthogonal. Therefore, reciprocal


vectors of the eigenbases do not differ from

p = p ,

and

p :

p = p .

The important problem for applications is to determine


given deformation gradient
method.

(3.24)
, p

and

by the

F. To solve the problem, one should use the following

31

1.3. Polar Decomposition

1) Construct the tensor

U2 = FT F (or V2 = F FT ) and find its components in

some basis being suitable for a considered problem; for example, in the Cartesian
basis

i :
e

2 )i j e
i e
j
U2 = (U

2) Find eigenvalues of the matrix

and

2 )i
(U
j

i e
j .
V2 = (V 2 )i j e

by solving the characteristic equation

det (U2 2 E) = 0,
which in the basis

i
e

(3.25)

takes the form

det ((U2 )i j 2 ji ) = 0.
3) Find eigenvectors

of the tensor

(3.25a)

and eigenvectors

of the tensor

from the equations

U2 p = 2 p ,
written, for example, in the basis

2 )i 2 i )Q
b j = 0,
((U
j
j

bj
Q

(3.26)

i :
e

where

V2 p = 2 p ,

b j = 0,
((V 2 )i j 2 ji )Q

(3.26a)

and

bj
Q

are Jacobian matrices of the eigenvectors:

bj e
j ,
p = Q

To determine the matrices

bj
Q

and

bj e
j .
p = Q

bj ,
Q

(3.27)

one should consider only independent

equations of the system (3.26a) and the normalization conditions (3.5):

|p | = 1,

|p | = 1,

(3.28)

which are equivalent to the quadratic equations

bi Q
b j ij = 1,
Q

bi Q
b j ij = 1.
Q

(3.28a)

4) Write the dyadic products (3.6) and find resolutions of the tensors
for the eigenbases; for example, for the Cartesian basis

U=

3
X

bi Q
bj e
i e
j ,
Q

V=

3
X

and

i :
e

bi Q
bj e
i e
j .
Q

=1

=1

Exercises 1.3.21.3.4 show examples of determination of the tensors

and

V.

Remark 2. Notice that a solution of the quadratic equations (3.28a) may be

i
b i , this
b
not unique due to the choice of signs of matrix components Q and Q

ambiguity is resolved by applying one more additional condition, namely the


condition of coincidence of the vectors

t 0+

and

p (t) = p (t),

when

= 1,

t 0+ :
2, 3.

32

Chapter 1. Kinematics of Continua

bi ,
Q

For the matrix

the ambiguity of the sign choice remains. However, if

p (x, t), then this ambiguity may be retained only


at one point x0 at one time, for example, t = 0; and for the remaining x and t,

b i is chosen from the continuity condition of the vector field p


a sign at Q
(x, t)

(for continuous motions). If the eigenvector field p (x0 , 0) contains the vectors

, then the remaining ambiguity is resolved by the condition p (x0 , 0) = e


.
e

there is a field of eigenvectors

The ambiguity of a solution of the system (3.26a), (3.28a) may also appear,

t1

if at some time

at a point

case, values of the matrices

the eigenvalues

b i (t1 )
Q

(t1 )

b i (t1 )
Q

and

prove to be triple. In this

are determined, as a rule, by

passage to the limit:

b i (t1 ) = lim Q
b i (t),
Q

b i (t1 ) = lim Q
b i (t), = 1, 2, 3.
Q
tt1

tt1

In the case of double eigenvalues

corresponding matrix components

bi
Q

these formulae are applied only to their

and

bi .
Q

1.3.3. Representation of the Deformation Tensors in Eigenbases.

Theorem 1.12. In the tensor bases p p and p p , the CauchyGreen


tensors

G, g1

J, the Almansi tensors A and ,


G1 , g have the diagonal form:

and

and

C=

3
X
1

=1

A=

(2 1)p p ,

3
X
1

=1

3
X
1

=1

and

G=

(1

)p

p ,

J=

=1
3
X

G1 =

2 p p ,

3
X

(3.29a)

(2 1)p p ;

p p ,

(3.29b)

=1

X
3

g1 =

2
(1
)p p ,

3
X
1

=1

2 p p ,

=1

and the deformation measures

g=

3
X

p p .

=1

On substituting formulae (3.6) into (3.22), we get (3.29a). Formulae (3.29b)

follow from (3.29a) and (2.7), (2.8).

Similarly to formulae (3.29), we can introduce new deformation tensors by


determining their components with respect to the bases
follows:

M=

3
X

=1
where
and if

f ( )p p ,

M=

3
X

p p

f ( )p p ,

or

p p

as

(3.30)

=1

f ( ) is a function of . If f (1) = 0, then we get the deformation tensors;


f (1) = 1, then we get the deformation measures.

33

1.3. Polar Decomposition

Among the tensors (3.30), the logarithmic deformation tensors and measures

3
X

H=

lg p p ,

H=

=1

3
X

lg p p ,

(3.31)

=1

H1 = H + E,

H1 = H + E,

are the most widely known; they are called the right and left Hencky tensors,
and also the right and left Hencky measures, respectively.
With the help of the eigenvectors
3
X

p p =

=1

3
X

and

p p O = (

=1

3
X

we can form the mixed dyads

p p ) O = E O.

(3.32)

=1

Here we have used the properties (3.23a) and (3.24), and the representation of
the unit tensor

in an arbitrary mixed dyadic basis.

Thus, the rotation tensor

accompanying the deformation can be expressed

in the eigenbasis as follows:

O=

3
X

p p = pi pi .

(3.33)

=1
On substituting (3.33) and (3.6) into (3.1) and taking (3.5) into account, we
get the following expression of the deformation gradient in the tensor eigenbasis:

F=OU=

3
X

p p

=1

3
3
X
X

p p .
p p =

=1

=1

According to (3.34), the transpose

(3.34)

FT

and inverse

F1

gradients are ex-

pressed as follows:

FT =

3
X

F1 =

p p ,

1.3.4. Geometrical Meaning of Eigenvalues.

Vectors of eigenbases

are connected by the transformation (3.23a). In

vectors

dx

oriented along the eigenbasis vectors

to radius-vectors

p p .

(3.35)

=1

=1

3
X

dx :

p ,

and

take elementary radius-

then in

dx = p |dx |,

dx = F dx .

they correspond

(3.36)

Substitution of (3.34) into (3.36) yields

dx =

3
X

=1

p p p |dx | = |dx |p ,

(3.37)

34

Chapter 1. Kinematics of Continua

dx in K will be
p .

vectors dx and dx by ds

i.e. the elementary radius-vectors

also oriented along the

corresponding eigenbasis vectors


Denote lengths of the

and

ds,

respectively, and

derive relations between them:

ds2 = dx dx = dx FT F dx = |dx |2 p FT F p =

= ds2 p G p = ds2 2 .

(3.38)

Here we have used equations (3.29b) and (3.36). Formula (3.38) proves the
following theorem.

Theorem 1.13.

Eigenvalues

(principal stretches) are the elongation ratios

for material fibres oriented along the principal (eigen-) directions:

= ds /ds .

(3.39)

1.3.5. Geometric Picture of Transformation of a Small Neighborhood of a

Point of a Continuum. In K, consider a small neighborhood of the material


0
point M contained in a continuum; then every point M , connected to M by the

elementary radius-vector

by radius-vector

dx

dx (Figure 1.12), will be connected to the same


K. These radius-vectors are related as follows:

point

in

dx = F dx.

(3.40)

The relation can be considered as the transformation of arbitrary radius-vector

dx

into

dx.

Figure 1.12. Transformation of a small neighborhood of the point contained in a continuum

Rewrite the relation (3.40) in Cartesian coordinates:

i
dxi = Fm
dxm ,
where

i
Fm

(3.41)

are components of the deformation gradient with respect to the

Cartesian basis (see Exercise 1.1.5):

i
Fm
= (xi / xm ),
which depend only on coordinates
of coordinates

dxm

of the point

of its neighboring points

M0 .

(3.42)
M,

but they are independent

Therefore the transformation

35

1.3. Polar Decomposition

(3.41) is a linear transformation of coordinates

dxm

dxi ,

into

i.e. this is an affine

transformation.
As follows from the general properties of affine transformations, straight

lines and planes contained in a small neighborhood in

K.

lines and planes in actual configuration

will be straight

Parallel straight lines and planes

are transformed into parallel straight lines and planes. Therefore if a small

neighborhood in

is chosen to be a parallelogram, then in

the neighborhood

will be a parallelogram as well (although angles between its edges, edge lengths
and orientation of planes in space may change).

(and, in general, a surface specified by an

n-th

order) is transformed into a surface of the

Since a second-order surface in


algebraic expression of arbitrary
same order in

K,

a small spherical neighborhood in

ellipsoid in actual configuration

As follows from formula (2.34), the ratio of lengths


vector (or of elementary radius-vector
initial length
of

dx

in

ds /ds

K)

and

of an arbitrary

is independent of the

of the vector (because the relative elongation

is independent

ds ).
According to the polar decomposition (3.1), the transformation (3.40) from

ds

is transformed into an

(Figure 1.12).

to

can always be represented as the superposition of two transformations:

dx = O dx0 ,

realized with the help of the stretch tensor

U,

(3.43)

and the rotation tensor

O,

dx = V dx0 ,
The stretch tensor

dx0 = U dx,

dx0 = O dx.

or

(3.44)

which has three eigendirections

p ,

transforms a small

M with compressing or extending the neighborhood

along these three directions p . The tensor O rotates the neighborhood deformed

along p as a rigid whole until the direction of p becomes the direction of


neighborhood of the point

p .

If one use the left stretch tensor

coincidence with

V,

U.
M

the eigendirection
point

in

till their

is first realized, and then compression or tension of the

neighborhood occurs along the direction


If a point

so rotation of axes

is connected to

The result will be the same as for

by radius-vector

dx

oriented along

(which is unknown before deformation), then in

will be connected to

corresponding eigendirection

by radius-vector

p .

If a small neighborhood of point


Figure 1.12), then in

p .

dx

the

oriented along the

is chosen to be a sphere in

(see

the sphere becomes an ellipsoid with principal axes

oriented along the eigendirections

p .

Thus, the transformation of a small neighborhood of every point

M contained

in a continuum under deformation can always be represented as a superposition

36

Chapter 1. Kinematics of Continua

of tension/compression along eigendirections and rotation of the neighborhood


as a rigid whole, and also displacement as a rigid whole.

Exercises for 1.3


Exercise 1.3.1.

and

Using the formula (3.21a), show that the following relations between

hold:

for all integer

Vm = O Um OT ,

Um = OT Vm O

(positive and negative).

Exercise 1.3.2.

Using the results of Exercises 1.1.1 and 1.2.1, show that for the
problem on tension of a beam, eigenvalues are

= k ,
U

The stretch tensors

and

= 1, 2, 3.

are coincident and have the form

U=V=

3
X

k e e ,

=1
and eigenvectors

and

coincide with

e :

p = p = e ,
The rotation tensor

Exercise 1.3.3.

= 1, 2, 3.
O = E.

for this problem is the unit one:

Using the results of Exercises 1.1.2, 1.2.2 and Remark 2, show that

for the problem on simple shear (see Example 1.2 from paragraph 1.1.1), the tensors
and

are expressed as follows:

2 )i j e
i e
j ,
U2 = FT F = E + aO3 + a2 e2i = (U

2 )i j
(U

i e
j ,
V2 = E + aO3 + a2 e2i = (V 2 )i j e
!

2
1
a
0
1+a
a
2 i
2

= a 1+a 0 ,
(V ) j =
a
1
0

eigenvalues

0
0

!
,

are

2 = 1 + b |a|, = 1, 2;
3 = 1,
q
q
a
a
b1 = + 1 + a2 /4 ,
b2 = 1 + a2 /4 ,
2

eigenvectors

and

p (a > 0)

are

p = p
1

p1 = q
1

the stretch tensors

+b

1
1

+ b2

1 + e
2 ),
(b1 e
2
1

and

2 ),
(
e1 + b e
1

p2 = q
1

+ b22

3 ,
p3 = e
1 + e
2 ),
(b2 e

are

ij e
i e
j = U0 e
21 + U1 O3 + U2 e
22 + e
23 ,
U=U

3 ,
p3 = e

U2

37

1.3. Polar Decomposition

i e
j = U2 e
21 + U1 O3 + U0 e
22 + e
23 ,
V = V ij e

!
U0 U1 0
U2 U1 0
ij = U1 U2 0 ,
U
V ij = U1 U0 0 ,
0

U =
and the rotation tensor

b1

+ b1 a

+ b21

b2

+ b2 a

+ b22

= 0, 1, 2,

has the form

2 e
2 e
1 ),
i e
j = cos (
22 ) + sin (
O = O je
e1 e
e21 + e

!
cos sin 0
i
j = sin cos 0 ,
O
b1

cos =
1

Show that functions

b1 (a)

+ b21

and

b1 + b2 = a,
Show that at

a=0

sin =

b2

b2 (a)

+ b22

b21
1

+ b21

b22

+ b22

satisfy the following relationships:

b1 b2 = 1,

b21 + b22 = 2 + a2 .

for the considered problem the following equations really hold:

b1 = 1,

b2 = 1,

1 = 2 = 3 = 1,

2 ),
e1 + e
p1 = p1 = (

2 ).
p2 = p2 = (
e1 e

Exercise 1.3.4.

Using the results of Exercise 1.2.3, show that for the problem on
rotation of a beam with tension (see Example 1.3 from paragraph 1.1.1), eigenvalues
have the form
= k ,
= 1, 2, 3,
and eigenvectors

,
p = e

,
p = O0 e

= 1, 2, 3.

Using formulae from Exercise 1.1.3 and data from Example 1.3, show that tensors
and also C, A, and J have the form

U, V, O,

U = U0 =

3
X

i e
j ,
O = O0 = O0i j e

e
,
k e

=1

21 + V1 O3 + V2 e
22 + k3 e
23 = V0 ij e
i e
j ,
V = O0 U0 OT
= V0 e
0

!
V0 V1 0
i
V0 j = V1 V2 0 ,
0
0
k3
V1 = k1 cos2 + k2 sin2 ,

V1 = (k1 k2 ) cos sin ,


2

V2 = k1 sin + k2 cos2 ,
1

C = (U20 E) =
2

3
X

=1

1
2

,
(k2 1)
e e

38

Chapter 1. Kinematics of Continua


3
X

2
= (E U
)=
0

A = (E V

=1

ij

ij

j ,
) = ( g )
ei e
2

gij

where metric matrices

and

g ij

1
2

,
(1 k2 )
e e
1

j ,
J = (V2 E) = (gij ij )
ei e

are determined by formulae from Exercise 1.1.3.

We should take into consideration that the tensors

and

do not feel the beam

rotation  they are coincident with the corresponding tensors for the problem on pure
tension of the beam. Show that if we change the sequence of transformations (i.e. we
first rotate and then extend the beam), then the tensors

and

do not feel the rotation.

1.4. Rate Characteristics of Continuum Motion


1.4.1. Velocity.

The velocity (vector) of the motion of a material point

with Lagrangian coordinates


radius-vector

x(X i , t)

Xi

with respect to time at fixed values of

v(X i , t) =
(X i , t)
t

Velocity components

vi

Xi

x
i ,
e
t

vi =

X i:

with respect to the basis

i =
v = vi e

is determined as the partial derivative of the

(4.1)
i
e

have the form

x
(X j , t).
t

(4.2)

1.4.2. Total Derivative of a Tensor with respect to Time. Any vector field

a(x, t)

(and also scalar or tensor field) varying with time, which describes

some physical process in a continuum, can be expressed in both Eulerian and


Lagrangian descriptions with the help of the motion law (1.3):

a(x, t) = a(x(X j , t), t).

(4.3)

Determine the derivative of the function with respect to time at fixed


for a fixed point

M):

respect to time

The partial derivative of a varying vector field


at fixed coordinates

function (4.3) with respect to time:

(i.e.

a
a
a xj
i=
i+ j
.
t X
t x
x t X i
Definition 1.1.

Xi

Xi

(4.4)
a

(4.3) with

is called the t o t a l derivative of the

a
da
=
.
dt
t X i

(4.5)

According to formulae (4.2), (1.11) and (1.23), the second summand on the
right-hand side of (4.4) can be rewritten as follows:

a xj
a
a
a
i e
j P kj
= vj P kj
= vi e
= v rk
= v a. (4.6)
j t
k
k
x
X
X
X k
Then the relationship (4.4) yields

a
da
=
+ v a,
dt
t

(4.7)

39

1.4. Rate Characteristics of Continuum Motion

where we have introduced the notation for the partial derivative with respect to
time which will be widely used below:

a
a i
=
(x , t) i .
t
t
x
In formula (4.7) the vector
that if

(4.8)

is considered as a function

a(xj , t).

It is evident

j
is considered as a function of (X , t), then from the definition (4.5) we

get

da i

(x , t) = a(X j , t) i .
dt
t
X

(da/dt)

The total derivative

(4.9)

is also called the material (substantial, indi-

(a/t) in
v a is the

vidual) derivative with respect to time,

(4.7) is the partial (local)

derivative with respect to time, and

convective derivative.

The material derivative


a fixed material point

M,

da/dt

in time at a fixed point

characterizes a change of the vector field

in

the local derivative determines a change of values of


in space, and from (4.6) we get that the convective

derivative characterizes a change of the field due to transfer of the material


particle

from a point

If we choose the vector

and the velocity

to a point x + vdt in space.


v as a, then the relationship between the displacement

vectors has the form

v=

dx
du
u
=
=
+ v u.
dt
dt
t

(4.10)

Similarly to formula (4.5), we can define the total derivative of the


tensor

n with respect to time:


n

n
= d n (xi , t) = (X i , t)

dt

Theorem 1.14.

Xi

nth-order
(4.11)

The total derivative (4.11) of a varying tensor field

n (xi , t)

can be written as a sum of local and convective derivatives:

d n
n
=
+ v n .
dt
t

(4.12)

Proof of the theorem is similar to the proof of the relationship (4.7). Details

are left as Exercise 1.4.6.

Let us consider now the question on components of the total derivative tensor.

Theorem 1.15.

Components of the total derivative tensor

with the corresponding components of a tensor


bases

i
ri , e

and

e
ri

and a moving basis

ri

as follows:

d i ...in

= i ...in (X i , t) i ,
dt
t
X
d i ...in
i ...in i
i ...in i

=
(x , t) + vk k
(x , t),
dt
t
x
e i ...in e i
d e i ...in
e k
e i ...in (X
e i , t),

=
(X , t) + vek
dt
t
1

are connected

with respect to stationary

(4.13)
(4.14)
(4.15)

40

Chapter 1. Kinematics of Continua

d i

dt

= i
t

in

1 ...

in

1 ...

n
X

(X , t) +

(i

1 ...

k...in

k v i )(X i , t),

(4.16)

=1

where

= d
i

in

1 ...

dt

in =
(xi , t)
ei . . . e
1

d ei

dt

1 ...

in

k in place of i .
H To prove the theorem,
i
=

1 ...

e i , t)e
rin =
(X
ri . . . e

In formula (4.16), the component

d i

dt

in

(X i , t)ri . . . rin =
1

d i

dt

1 ...

in

(X i , t)ri . . . rin .
1

...k...in

as the

th

superscript has index

n for different bases:

we resolve the tensor

in = i ...in (X i , t)ri . . . rin =


(xi , t)
ei . . . e
e i ...in (X
e i , t)e
=
ri . . . e
rin = i ...in (X i , t)ri . . . rin ,

...in

(4.17)

(4.18)

n
and choose arguments of components of the tensor as in formula (4.18).

Then, substituting the resolution (4.18) for the basis ri into the definition

(4.11), we get the expression (4.13), because dri /dt = 0.


i into the relationship
On substituting the resolution (4.18) for the basis e
(4.12), we obtain

i ...in
i
=
i . . . e
in + vk e
k m

e
1

...in m

e
i . . . e
in . (4.18a)
e

It is evident that formula (4.14) follows from (4.18a).


In a similar way, substituting the resolution (4.18) for the basis
relation (4.12) and using the property (1.61) of nabla-operators
also the equation

e
ri /t = 0

(for the stationary basis

e
ri

e
ri

and

into the

e,

and

see paragraph 1.1.7), we

get

e i ...in
=
e m
ei
e

ri . . . e
rin + veke
rk
1

...in m

e
r e
ri . . . e
rin ;
1

(4.19)

and formula (4.15) follows from (4.19) at once.


Finally, substituting the resolution (4.18) for the moving basis

ri

into the

definition of the total derivative (4.12), we obtain

i ...in

ri . . . rin +
i

n
X

...i ...in

ri . . .
1

=1

ri
. . . rin .
t X i

(4.20)

Due to the definition (1.10) of local bases vectors and the definition (4.1) of the
velocity, we have

2x
ri
v

(X j , t) j =
j =
= i v k rk .
i

t
X
tX X
X i X j

(4.21)

On substituting (4.21) into (4.20) and then collecting components at the same
elements of the polyadic basis, we derive the formula (4.16).

41

1.4. Rate Characteristics of Continuum Motion

It should be noted that arguments of the resolutions (4.18) and of the


derivatives of tensor components (4.13)(4.16) have been chosen in the specific
way.

1.4.3. Differential of a Tensor.


Definition 1.2. For a tensor field n (xi , t),

d n =

the following object

d
dt
dt

(4.22)

is called the d i f f e r e n t i a l of a tensor field (or the differential of a tensor)

n (xi , t).

According to formula (4.12) for the total derivative of a tensor with respect
to time, we get that the differential of a tensor can be written in the form

d n (xi , t) =

n
t

+ v n dt.

(4.23)

According to (4.10), the relation (4.23) takes the form

d n =

n
dt + dx n
t

When a tensor field is stationary (i.e.


tensor field has the form

(4.24)

n /t = 0),

db n = dx n .

For stationary tensor fields

db n = d n ,

(4.25)

but in general these differentials

are not coincident.


According to Theorem 1.15, components of the tensor
the fixed basis

ri

d n

with respect to

are written as follows:

the differential of the

d = d

j1 ...jn

rj . . . rjn ,
1

j1 ...jn

d j ...jn
=
dt.
dt
1

(4.26)

From (4.22) and (4.7) we get the following expression for the differential of
a vector:

da(X i , t) =

da
dt =
dt

a
t

+ v a dt,

(4.27)

and from (4.25) we have

db
a = (v a)dt = ( a)T dx.
In particular, if

a = x,

(4.28)

then, by formulae (4.28) and (1.35a), we obtain

dx = ( x)T dx = F1 dx,

(4.29)

b
dx = F dx.

(4.30)

or

On comparing formulae (4.30) with (1.34), we find that the elementary radiusvector

dx,

introduced in paragraph 1.1 and connecting two infinitesimally close

material points

and

M0 ,

coincides with the vector

b
dx

in the notation (4.25).

42

Chapter 1. Kinematics of Continua

1.4.4. Properties of Derivatives with respect to Time. Let us establish now


important properties of partial and total derivatives of vector fields with respect
to time.

Theorem 1.16.

The partial derivative of the vector product of basis vectors

with respect to time has the form

r
r
(r r ) = r + r .
t
t
t

(4.31)

Determine the derivative of the vector product of two local basis vectors with

respect to time:

i Qj e
j ) = (Qi Qj )
j =
(r r ) = (Qi e
ei e
t
t
t
Qj
Qi
j
i
i Q e
j + Q e
i
j .
=
e
e
t
t
With use of relation (1.10) we really get (4.31).

Theorem 1.17. For arbitrary continuously differentiable vector


=a
i (xk , t)
ei and b(x, t) = bi (xk , t)
ei , we have the formulae

fields

a(x, t) =

a
b

(a b) =
b+a
,
t
t
t

a
b
(a b) =
b+a
,
t
t
t

a
b
(a b) =
b+a
.
t
t
t

A proof is similar to the proof of Theorem 1.16.

Theorem 1.18.

(4.33)
(4.34)

The total derivatives of the vector and scalar products of two

arbitrary vector fields

a(x, t)

and

b(x, t)

with respect to time have the forms

da
db
d
(a b) =
b+a ,
dt
dt
dt
d
da
db
(a b) =
b+a .
dt
dt
dt

(4.32)

(4.35)
(4.36)

To prove formula (4.35), one should use the property of the total derivative

(4.7):

(a b) = (a b) + v (a b).
dt
t

Modify the first summand by formula (4.32) and the second summand  by the
formula

(a b) = ( a) b ( b) a

[12], then we get

a
b
d
(a b) =
b
a + v ( a) b v ( b) a.
dt
t
t
Collecting the first summand with the third one and the second summand with
the fourth one, and using the property (4.7) of the total derivative of a vector,
we obtain

d
da
db
da
db
(a b) =
b
a=
b+a .
dt
dt
dt
dt
dt

Formula (4.36) can be proved in a similar way.

43

1.4. Rate Characteristics of Continuum Motion

Theorem 1.19.

The total derivative of

the divergence of the velocity

with respect to time is connected to

by

d
g = g i v i = g v.
dt
H

(4.37)

Let us differentiate the second relation of (1.15) with taking formula (4.31)

into account:

d
2x
d
g = r1 (r2 r3 ) =
(r2 r3 )+
dt
dt
tX 1

2x
2x
r3 + r1 r2
.
+ r1
tX 2
tX 3
Since

(4.38)

2x
v
=
= i v = i v j rj ,
i
tX
X i

we get

d
g = 1 v g r1 + r1 (2 v r3 ) + r1 (r2 3 v).
dt

(4.39)

Here we have used the relations from Exercise 1.1.14.


According to the definition of the vector product (0.2), we obtain

r1 2 v r3 = r1

g ijk 2 v i 3j rk = g i31 2 v i = g 2 v 2 .

On substituting (4.40) into (4.39), we really get formula (4.37).

(4.40)

1.4.5. The Velocity Gradient, the Deformation Rate Tensor and the Vor
ticity Tensor. Consider elementary radius-vectors dx and dx connecting two

0
infinitesimally close points M and M in configurations K and K, respectively.
0
Determine the velocity of the point M relative to the configuration connected

b :
dv

v T

2x
2x
i
i
b
dv = dx =
dX =
r dx = r
dx = v dx.
i
i
i

to the point

M.

To do this, determine the velocity differential

X t

X t

(4.41)

Here we have used the second equation of (1.33), the definition of the gradient
(1.24) and formula (4.1). In a similar way, using the first equation of (1.33):

dX i = ri dx,

we get one more expression for the vector

b :
dv

b = ( v)T dx.
dv
The second-order tensor
connects the relative velocity

dx

itself:

( v) is called
b of an elementary
dv

b = L dx,
dv

(4.42)
the velocity gradient, which
radius-vector

to the vector

L = ( v)T .

(4.43)

can be represented by a

Just as any second-order tensor (see [12]), the tensor


sum of the symmetric tensor

dx

and the skew-symmetric tensor

L = D + W.

W:
(4.44)

44

Chapter 1. Kinematics of Continua

The symmetric deformation rate tensor

is determined as follows:

D = ( v + vT ).

(4.45)

This tensor has six independent components.


The skew-symmetric vorticity tensor

is determined as follows:

W = ( vT v).

(4.46)

Since the tensor

is skew-symmetric and has three independent components,

we can put the tensor

in correspondence with the vorticity vector

connected to the tensor (see [12]) as follows:


1

= W ,

W = E.

where

(4.47)

is the Levi-Civita tensor, which has the third order (see [12]). This

tensor is determined as follows:


1

= ijk ri rj rk .

(4.48)

On substituting (4.44)(4.47) into (4.42), we prove the following theorem.

Theorem 1.20 (CauchyHelmholtz). The velocity v(M0 ) of an arbitrary point


M0 in a neighborhood of the material point M consists of the translational
motion velocity

v(M)

of the point

M, the velocity dx
D dx, i.e.

of rotation as a

rigid whole and the deformation rate

or

Example 1.4.

b = dx + D dx
dv

(4.49)

v(M0 ) = v(M) + dx + D dx + o (|dx|).

(4.49a)

Determine the tensor

for the problem on tension of a beam

(see Example 1.1), substituting (4.2) into (4.43):


3
3
X

X
i

e
= L.


=
L =e
v =e
k X e
k e
X i
X i

=1

Since the velocity gradient

=1

in this case proves to be a symmetric tensor, from

(4.45) and (4.46) it follows that

D = L,
Thus, in this case

Example 1.5.

W = 0.

= 0.

Determine the tensor

for the problem on simple shear (see

Example 1.2), substituting formula (4.2) into (4.43):

i
LT = e

vj i

v1

2 e
1 = a
1 .
=
e

e
=
e
e2 e
j
X i
X i
X 2

According to formulae (4.45) and (4.46), we get

2 + e
2 e
1 ),
D = (a/
2)(
e1 e

45

1.4. Rate Characteristics of Continuum Motion

2 e
2 e
1 ) = (a/
j .
W = (a/
2)(
e1 e
2)(1i 2j 2i 1j )
ei e
Using formula (4.47), we determine the vorticity vector
1

k = (21k 12k )
3 ,
ek = e
= W = (1i 2j 2i 1j )jik e

which is orthogonal to the shear plane.

1.4.6. Eigenvalues of the Deformation Rate Tensor. Just as any symmetric

tensor, the deformation rate tensor

has three orthonormal real-valued eigen-

vectors and three real positive eigenvalues (see [12]). Denote the eigenvectors
by

 by

(these vectors, in general, are not coincident with

D .

Then the tensor

D=

p )

and the eigenvalues

can be resolved for its dyadic eigenbasis as follows:

3
X

D q q ,

q q = .

(4.50)

=1

K an elementary radius-vector dx , connectM and M0 , so that the vector is oriented along the eigenvector q of
tensor D; then, similarly to (3.36), we can write

Take in the actual configuration


ing points
the

|dx | = (dx dx )1/2 .

dx = q |dx |,

(4.51)

Apply the CauchyHelmholtz theorem (4.49) to the elementary radius-vector:

b = dx + D dx .
dv

(4.52)

dx
dx ( dx ) = 0,

Multiplying the left and right sides of the equation by

and taking account

of the property of the mixed derivative

we get

b = dx D dx .
dx dv
Substituting in place of

(4.53)

its expression (4.50) and in place of

dx

their

expressions (4.51), we obtain

b = |dx |2
dx dv

3
X

D q q q q = D |q |2 .

(4.54)

=1
Here we have used the property (4.50) of orthonormal vectors

q .

Modify the scalar product on the left-hand side as follows:

b = dx dx =
dx dv
t

(dx dx ) = |dx | |dx |.


2 t
t
1

(4.55)

On comparing (4.54) with (4.55), we obtain the following theorem.

Theorem 1.21.

Eigenvalues

of

the

deformation

rate

tensor

are

the

rates of relative elongations of elementary material fibres oriented along the


eigenvectors

q :
D =

|dx |.
|dx | t
1

(4.56)

46

Chapter 1. Kinematics of Continua

1.4.7. Resolution of the Vorticity Tensor for the Eigenbasis of the Deformation Rate Tensor. Modify the right-hand side of (4.52) as follows:

b = dx + D dx = ( q + D q )|dx |,
dv

(4.57)

and the left-hand side of (4.52) with taking (4.56) into account:

b = dx = (|dx |q ) = |dx | q + |dx | q = |dx | D q + q .


dv
t
t
t
t
t
(4.58)
On comparing (4.57) with (4.58), we get the following theorem.

Theorem 1.22.

The vorticity tensor

rate of changing the eigenvectors

q =

(or the vorticity vector

to the vectors

connects the

themselves:

q
= q = W q .
t

Using formula (4.59), we can resolve the tensor

(4.59)
for the eigenbasis

of

the deformation rate tensor as follows:

W=

3
X

q q = q i qi .

(4.60)

=1

1.4.8. Geometric Picture of Infinitesimal Transformation of a Small Neighborhood of a Point. If in configuration K at time t we consider an elementary
0
radius-vector dx connecting two infinitesimally close material points M and M ,

dt the radius-vector
K(t + dt) (Figure 1.13):

then for infinitesimal time

dx0

in configuration

dx = x0 (t) x(t),

dx0 = x0 (t + dt) x(t + dt),

(4.61)

M and M0 in configuration
0
K(t), respectively; and x(t + dt) and x (t + dt)  in configuration K(t + dt).
0
Displacements of points M and M for infinitesimal time are defined by the
0
velocity vectors v(M) and v(M ), respectively:

where

x(t)

and

x0 (t)

is transformed into radius-vector

are radius-vectors of the points

x(t + dt) x(t) = v(M) dt,

x0 (t + dt) x0 (t) = v(M0 ) dt.

Figure 1.13. Infinitesimal transformation of an elementary radius-vector

(4.62)

47

1.4. Rate Characteristics of Continuum Motion

Formulae (4.61) and (4.62) and simple geometric relations (see Figure 1.13)
give

v(M0 )dt v(M)dt = dx0 dx.

(4.63)

On substituting (4.63) into (4.49a), we obtain the relation between elementary


radius-vectors

dx0

and

dx:

dx = dx + dt dx + dtD dx + dt o(|dx|).

(4.64)

The relation (4.64) can be considered as the transformation of coordinates

dxi dx0i in a small neighborhood of the point contained in a continuum. Since


dt and dtD are independent of dx and dx0 , so the transformation is linear,

i.e. affine. The relation (4.64) can be represented as a superposition of two


transformations up to an accuracy of

o(|dx|):

00

dx = AD dx,
0

00

dx = Q dx
AD

The tensor

AD = E + dtD,

(4.65)

Q = E + dt E.

(4.66)

is symmetric and has three eigendirections, which are

coincident with the eigendirections


So just as the tensor

U,

the tensor

of the deformation rate tensor

AD

D.

transforms a small neighborhood of

M by extending or compressing the neighborhood along the principal


q . The material segments |dx00 | oriented along the eigendirections q

a point

directions

retain their orientation under the transformations (4.65), but their lengths vary
as follows:

dx00 = (1 + D dt|dx |)q = (1 + D dt)dx .

dx = |dx |q ,
The tensor
because

(4.66) is orthogonal up to an accuracy of values

T
2
2
Q QT
= (E + dtW) (E + dtW ) = E (dt) W .

Here we have taken into account that the vorticity tensor

M-point

neighborhood as a rigid whole for infinitesimal time

The vorticity vector

forming the tensor

(4.67)

is skew-symmetric.

Thus, the transformation (4.66) determined by the tensor


the

(dt)2 ,

is rotation of

dt.

can be considered as instanta-

neous angular rate of rotation of the small neighborhood as a rigid whole, or as


instantaneous angular rate of rotation of the eigentrihedron
rate tensor relative to the fixed basis

i .
e

of the deformation

This fact will be considered in detail in

paragraph 1.5.7.
On uniting the properties of the transformations (4.65) and (4.66), we can
make the following conclusion.

Theorem 1.23.

The infinitesimal transformation of a small neighborhood of

the point contained in a continuum is a superposition of tension-compression


of the neighborhood along the eigendirections

and rotation of the axes

as a rigid whole about the axis with the direction vector

Thus, we have the certain analogy between the eigendirections

of the

q of the tensor D: elementary material fibres

oriented along p and along q remain mutually orthogonal and undergo only

tension-compression. The axes p remain mutually orthogonal under any finite

tensor

and the directions

48

Chapter 1. Kinematics of Continua

K
K(t + dt).

transformations from
from

K(t)

to

to

K,

but

 only under infinitesimal transformations

1.4.9. Kinematic Meaning of the Vorticity Vector.


tensor, the orthogonal tensor

Just as any orthogonal

of infinitesimal rotation from

K(t)

K(t + dt)

to

can be represented in the form (see [12])

Q = E cos(d) + (1 cos(d))e e e E sin(d),


where

is the infinitesimal angle of rotation of the trihedron

with the direction vector

e.

Since values of

(4.68)

about the axis

are infinitesimal, we have

Q = E e Ed.

(4.69)

On comparing (4.69) with (4.66), we get

=
i.e. the vorticity vector

e,

and the length

trihedron

||

d
e,
dt

|| =

d
,
dt

(4.70)

is really oriented along the instantaneous rotation axis

is equal to the instantaneous angular rate of rotation of the

of the deformation rate tensor.

Let us consider now the question: relative to what system the vorticity vector

defines the rotation rate.


To answer the question, we introduce another orthogonal rotation tensor

i ,
OW = qi e
which transforms the Cartesian trihedron
mal trihedron

The tensor

qi :

OW

i
e

(4.71)

as a rigid whole into the orthonor-

i .
qi = OW e

is a function of time

t,

because

(4.72)
qi = qi (t).
W takes

According to (4.60) and (4.72), the vorticity tensor

the form

W OT .
W = q i qi = O
W

(4.73)

With the help of (4.73) we can represented the orthogonal tensor


follows:

Thus, at each time


borhoods of a point

W OT .
Q = E + dtW = E + dtO
W
t

two orthogonal tensors

in the reference

OW

and

as

(4.74)
connect local neigh-

and actual configurations

K(t + dt).

00
If in K we consider an elementary radius-vector dx , then in K we find its
00
corresponding radius-vector dx obtained with the help of the rotation tensor
OW , and in K(t + dt)  radius-vector dx0 :

dx00 = OW dx00 ,

dx0 = Q dx00 .

A fixed observer connected to the Cartesian trihedron


formations: finite rotation for time

t,

instantaneous rotation of a local neighborhood for time


the infinitesimal rotation tensor

i
e

sees both the trans-

which is described by the tensor

dt,

OW ,

and

which is described by

49

1.4. Rate Characteristics of Continuum Motion

Thus, the vorticity vector


rotation of the trihedron

is the vector of instantaneous angular rate of

relative to the trihedron

i .
e

Comparing (4.66) with (4.74) (or (4.73) with (4.47)), we get

W OT = E.
O
W

(4.75)

1.4.10. Tensor of Angular Rate of Rotation (Spin). In paragraph 1.3 we


W OT , where OW is the orthogonal tensor of
O
W
QT can be set up for any orthogonal tensor Q
rotation. Such tensor = Q

have introduced the tensor

t.
QT is skew-symmetric, because
Q
QT + Q Q
T = (Q QT ) = (E) = 0,
Q

depending on time
The tensor

i.e.

(4.76)

QT )T = Q Q
T = Q
QT = .
T = (Q

(4.77)

. This tensor characterizes the angular rate of rotation of the orthonormal


trihedron

hi

formed with the help of

Q:

i ,
hi = Q e
relative to the Cartesian trihedron

(4.78)

i .
e

QT we can
Q
QT ,
Q = E + dtQ

Indeed, with the help of the tensor

form the tensor (4.74):

(4.79)

which, according to (4.67), is the orthogonal tensor of infinitesimal rotation; and


this tensor can be represented in the form (4.68) or (4.69):

Q = E de E,
where

(4.80)

is the infinitesimal angle of rotation of the trihedron

with the direction vector

e.

hi

about the axis

Comparing (4.79) with (4.80), we get the expression

QT = d e E,
Q

(4.81)

dt

which makes clear the sentence that the tensor

QT
Q

characterizes the instan-

d/dt of rotation of the trihedron hi about


QT is called the tensor of angular rate of
Q

taneous angular rate


The tensor

the axis

spin.

Expressing the tensor

= h , = 1, 2, 3,

from (4.78) in terms of the bases

e.

rotation or the

hi

and

i (h =
e

as the vectors are orthonormal):

i ,
Q = hi e

(4.82)

we get another representation of the spin:

QT = h i hi .
Q
Thus, we have proved the following theorem.

Theorem 1.24.

The spin connects the rates

h i

(4.83)
and the vectors

hi

defined by

formula (4.78) as follows:

QT ) hi .
h i = (Q

(4.84)

50

Chapter 1. Kinematics of Continua

Since the spin

QT
Q

is a skew-symmetric tensor, we can introduce the

corresponding vorticity vector

h:
QT = h E
Q

(from formulae (4.81) and (4.85) it follows that


(4.84) takes the form

Resolving the vector

(4.85)

h = (d/dt)e);

then formula

h i = h hi .
h

(4.86)

for the orthonormal basis

hi : h = jh hj ,

we get one

more representation of formula (4.84):

h i = jh hj hi = jik jh hk .

(4.87)

This formula can also be rewritten in the form

h = h h ,

6= 6= 6= .

(4.88)

Taking different orthogonal tensors (or orthonormal bases) as

(or

hi ),

we

obtain different spins.


1) If we choose eigenvectors of the stretch tensor

according to (4.83), the corresponding spin

i
U OT = p
U = O
U
i p ,

as

hi ,

i.e.

hi = pi ,

then,

takes the form

i ,
OU = pi e

(4.89)

and formulae (4.84) yields

pi = U pi .
2) If

hi = pi ,

then the corresponding spin

(4.90)

and the rotation tensor

OV

have the forms

V OT = p i pi ,
i ,
V = O
OV = pi e
V
p i = V pi .
3) If

hi = qi , then the corresponding spin W coincides


W (see (4.73)):
W OT = q i qi = W,
W = O

(4.91)
(4.92)

with the vorticity

tensor

(4.93)
(4.94)

q i = W qi .

O accompanying deformation as Q, then,

O connects two moving bases pi and pi :

4) If we take the rotation tensor


shown in (3.23a), the tensor

pi = O pi .
The tensor

can be expressed in terms of

(4.95)
OV

and

OU

as follows:

i e
j pj = OV OT
O = pi pi = pi e
U.
The corresponding spin

as

(4.96)

has the form

OT = (O
V OT + OV O
T ) OU OT =
=O
U
U
V
T
T
T

= OV OV + OV OU OU OV = V OV U OT
V.

(4.97)

51

1.4. Rate Characteristics of Continuum Motion

Unlike the cases 1)3), the spin tensor


rotation of the trihedron
to the trihedron

i
e

pi

characterizes the angular rate of

relative to the moving trihedron

pi ,

but not relative

being fixed.

Therefore, for the cases 1)3) the spins characterize the total angular rate,
and for the case 4)  the relative rate.

1.4.11. Relationships between Rates of Deformation Tensors and Velocity


Gradients. In continuum mechanics, one often needs the relations between
rates of the deformation tensors (and also measures) and the velocity gradients

L = ( v)T

and

L = ( v)T .

(4.98)

Let us derive these relations.

Theorem 1.25.

(F1 )

The rates of varying the gradient

are connected to

and

and the inverse gradient

by the relations

= L F,
F
(F1 ) = F1 L,
H

= L,
F

(F1 ) = F1 F1T L.

(4.99)

and taking the

= ( v)T F.
F

(4.99a)

Differentiating the relationships (1.35a) with respect to time

definition of the velocity (4.1) into account, we get

T = v = FT v,
F
According to the definitions of tensors
obtain formulae (4.99).
Differentiating the identity

= F

(F1 ) ; whence we get

(4.43) and

=
(F F1 ) = E

(4.98), from (4.99a) we

0, we find that

F F1 =

F1 .
(F1 ) = F1 F

(4.100)

On substituting the first two formulae of (4.99) into (4.100), we obtain

(F1 ) = F1 ( v)T ,

(F1T ) = ( v) F1T ,

i.e. the third and the fourth relationships of (4.99) hold as well.

(4.101)

According to formulae (4.45), (4.99a) and (4.101), we find that the rate of
the deformation gradient is connected to the deformation rate tensor

by the

relations
1
F1 + F1T F
T ),
D = (F
2

1
1 + F
1T FT ).
D = (F F
2

(4.102)

1 (F1 ) .
F
,A
,
, J and deformaTheorem 1.26. The rates of the deformation tensors C

, (G ) and (g ) are connected to the velocity gradients


tion measures G, g
Here and below we will use the notation

and

by the relationships

= FT D F,
C

= F1 D F1T ,

= D LT A A L,
A
J = D + L J + J LT ,

(4.103)

52

Chapter 1. Kinematics of Continua

and

= 2FT D F,
G
(G1 ) = 2F1 D F1T ,

g = LT g g L,
(g1 ) = L g1 + g1 LT ;

(4.104)

T
T

J = (1/2)(L FT + FT L),

C = (1/2)(F L + L F),

= (1/2)((E 2A) L F1 + F1T LT (E 2A)),


A

= (1/2)(F1 L (E 2) + (E 2) LT F1T ),

(4.105)

T
T

,
g = (g L F1 + F1T LT g),

G = F L + L F

1 ) = (F1 L G1 + G1 LT F1T ),
(G

1
(g ) = L FT + FT L.

(4.106)

and also

and

To prove formula (4.103), we must differentiate the relationships (2.5) with

respect to

and apply formulae (4.99):

= 1 (F
T F + FT F)
= 1 (FT LT F + FT L F) =
C
2

= FT (LT + L) F = FT D F,
2

(4.107)

1 ) =
= 1 (F 1T F1 + F1T F
A
2

= (LT F1T F1 + F1T F1 L) =


2

= (L (E 2A) + (E 2A) L) = D LT A A L,
2

(4.108)

1 F1T + F1 F
1T ) = 1 (F1 L F1T +
= 1 (F

+F

1T

L F

)= F
2

(L + LT ) F1T = F1 D F1T ,

(4.109)

1
1
J = (F FT + F F T ) = (L F FT + F FT LT ) =
2

= (L (E + 2J) + (E + 2J) LT ) = D + L J + J LT .
2

(4.110)

Formulae (4.104) follow from (4.103), if we have used the connections (2.7) and
(2.8) between the deformation tensors and measures.
Formulae (4.105) follow from (4.107)(4.110), if we have gone from

L = L F1 .

to

L:

(4.111)

53

1.4. Rate Characteristics of Continuum Motion

Using the connections (2.7) and (2.8) between the deformation tensors and

measures, from (4.105) we get formulae (4.106).

,V

Relations between the tensors

and the velocity gradients

and

more complicated. To derive them, we should use representations of

and

are

in

terms of the eigenbasis vectors.

Theorem 1.27.

The following expressions for the velocity gradients hold:

3
X


L =v = x=
p p + U FT FT V ,
t

(4.112)

=1

3
X

L = ( v) =

,=1


+ U

where

are components of the tensor

p p + V ,

(4.113)

with respect to the eigenbasis

p :

U = pi pi =

3
X

U p p ,

U = p U p = p p . (4.114)

=1

To prove formula (4.112), we should consider the first formula of (4.99a) and

substitute into this formula the expression (3.35) for

T =
LT = F

3
X

in the eigenbasis:

( p p + (p p + p p )).

(4.115)

=1
Using formulae (4.90) and (4.92), from (4.115) we derive the relationship
(4.112).
To prove formula (4.113), we use formulae (4.112) and (4.111), having
expressed

F1

in the form (3.35); then we get

L = L F1 = (

3
X

p p + (p p + p p ))

=1

3

X

,=1

3
X

p p =

=1

p p +

X

p p .
(p p )p p +

Here we have taken into account that the vectors

are orthonormal. Using

formulae (4.90) and (4.92), from (4.116) we derive the formula (4.113).
From formula (4.113) it follows that the deformation rate tensor
vorticity tensor can be represented in terms of the eigenbasis

D=

3
X

,=1

(4.116)

=1

1
2

N
and the

as follows:

U p p ,

(4.117)

54

Chapter 1. Kinematics of Continua

W=

1
2

3
X

,=1

U p p + V .

Here we have taken into account that the tensors

and

(4.118)
V

are skew-

symmetric.
Denote components of the tensor

D=

3
X

with respect to the basis

D p p ,

by

D :

D = p D p .

(4.119)

,=1
Then from (4.117) and (4.119) we get that diagonal components of the deformation rate tensor

with respect to the eigenbasis

determine the

relative rates of lengthening the material fibres oriented along the eigenvectors

(compare with formula (4.56)):

D = / = ds /ds ,
D are connected
2
2
1
=
U ,

and off-diagonal components

= 1, 3;

to

(4.120)

by the relations

6= .

(4.121)

From formulae (4.119) and (4.121) we can express the components

in

terms of components of the deformation rate tensor:

U =

D ,

6= ;

U = 0.

The diagonal components

are equal to zero, because the tensor

(4.122)
U

is

skew-symmetric.
On substituting the relationships (4.122) into (4.118), we find the expression

V of the tensor V with respect to the basis p in terms


W and D (and, hence, in terms of the velocity gradient L):

for components
the tensors

V = p V p = p W p
where

V =

3
X

2 + 2
2 2

D ,

V p p .

6= ,

of

(4.123)

(4.124)

,=1

Remark. The expressions (4.122) and (4.123) are valid only if the eigenvalues

6= ( 6= ; , = 1, 2, 3). If within the interval [t1 , t2 ] all


= ( = 1, 2, 3), then the stretch tensors

i
are spherical: U = pi p = E, V = E, and the eigenbases are not uniquely

defined: as pi and pi we can take any orthonormal triple of vectors. In particular,

one of the bases can be taken as fixed t [t1 , t2 ], for example, pi can be chosen
are not multiple:

three eigenvalues are coincident:

55

1.4. Rate Characteristics of Continuum Motion

case pi

pi (t1 ); and the second basis pi can


t [t1 , t2 ], and from (4.114) and (4.118)

as coincident with

depend on time

t.

In this

it follows that within the

considered time interval:

U = 0,
V = W,

U = 0,

V = p W p ,

(4.125)
, = 1, 2, 3.

These relationships take the place of formulae (4.122), (4.123) in this case.

[t1 , t2 ] only two of three eigenvalues are coincident,

for example, = , then their corresponding eigenvectors p and p are not

uniquely defined as well: only their orthogonality to the vector p , corresponding

to the third eigenvalue , is given. Then we can extend the definition of p and

p so that p p = 0 t [t1 , t2 ]. In this case it follows from (4.114) that the


If within the time interval

only component

vanishes, but

U 6= 0

and

It follows from (4.118) that the component


formula

V = p W p ,

U = 0.

U , U

The remaining components

U 6= 0.
V is determined

and

V , V

by the

(4.126)
are determined by

formulae (4.122) and (4.123).


If the situation with multiple roots appears only at some time

values of

U (t)

and

V (t)

t,

then the

can be determined by passing to the limit.

Substituting formulae (4.114) and (4.118) into equation (4.97), and taking
expressions (4.89) and (4.91) for
representation of the spin

=W

1
2

3
X
2 + 2

,=1

OU

OV
p :

and

in the basis

into consideration, we obtain the

U p p

3
X

i ) p p (
U (pi e
ej pj ).

(4.127)

,=1
Introducing the notation for direction cosines

,
= p e

,
l = p e

(4.128)

substituting (4.122) into (4.127) and collecting like terms, we obtain the following expression of the spin

in terms of

and

(i.e. in terms of

e,
=W+
e =

3
X

,=1

e p p ,

e =

L):
(4.129)

3
X

,=1

e ,

56

Chapter 1. Kinematics of Continua

3
X

e =

,=1

Theorem 1.28.

2 2

((2 + 2 ) 2 l l )D .

Rates of the deformation measures

are connected to the velocity gradient

, (U1 )
U

and

, (V1 )
V

by the formulae

e O + OT (D +
e T ) F),
= 1 (FT (D + )
U
2

1
e O + OT (D
e T ) F1T ),
) = (F1 (D )

(U

= 1 ((L + ) V + V (LT + T )),


V
2

(4.130)

(V1 ) = (( LT ) V1 + V1 (T L)).
2

Let us express the tensors

and

U = OT F,
Since

and

from the polar decomposition (3.1):

V = F OT .

(4.131)

are symmetric tensors, these expressions can be rewritten in the

symmetrized form
1

U = (FT O + OT F),
2

V = (F OT + O FT ).
2

(4.132)

Let us differentiate these relationships:

= 1 (F
T O + FT O
+O
T F + OT F)
,
U
2

OT + F O
T+O
FT + O F
T ).
= 1 (F
V
2

On substituting formulae (4.99) and the expression (4.97) for the spin

(4.133)

into

(4.133), we obtain

= 1 (FT LT O + FT O
OT O + OT O O
T F + OT L F) =
U
2

= (FT (LT + ) O + OT (L + T ) F),


2

(4.134)

= 1 (L F OT + F OT O O
T+O
OT O FT + O FT LT ) =
V
2

= ((L + ) V + V (LT + T )).


2

Taking formulae (4.129) and (4.44) into consideration, we find that

e = D + .
e
LT + = D + WT + W +

(4.135)

Substituting (4.135) into (4.134), we really get the first and the third formulae
of (4.130).
The remaining two formulae in (4.130) can be proved in a similar way.

57

1.4. Rate Characteristics of Continuum Motion

In continuum mechanics the deformation tensors

B and Y

are applied, which

have no explicit expression; they are defined by their derivatives and initial
values:

= 1 (U
U1 + U1 U)
,
B

B(0) = 0,

(4.136)

= 1 (V
V1 + V1 V)
,
Y

Y(0) = 0.

(4.137)

After substitution of the expressions (4.131), formula (4.136) takes the form

T F + OT F)
U1 + U1 (F
T O + FT O))
=
= 1 ((O
B
2

= (OT ( + L) F U1 + U1 FT (LT + T ) O) =
2

= (OT ( + L) O + OT (LT + T ) O),


2

(4.138)

and formula (4.137) is rewritten as follows:

= 1 ((F
OT + F O
T ) V1 + V1 (O
FT + O F
T )) =
Y
2

= (L F OT V1 + F OT T V1 + V1 O FT +
2

+V

O FT LT ) = (L + V V1 + V1 V + LT ).
2

Finally, we get the following expressions for

and

(4.139)

:
Y

= OT D O,
B

(4.140)

= D + 1 (V T V1 + V1 V).
Y

(4.141)

1.4.12. Trajectory of a Material Point, Streamline and Vortex Line. HavX i of a material point M in the motion law (1.3), we get

ing fixed coordinates

the parametric equation of a certain curve, where time

xi = xi (X k , t),

is a parameter:

6 t 6 t0 .

(4.142)

t = 0 is a point with

xi (X k ) of the material

The origin of the curve at


Cartesian coordinates

M in K, and the end of the curve at t = t0


i
k 0
is a point with Cartesian coordinates x (X , t )
0
of the point M in K(t ) (Figure 1.14). The
point

curve (4.142) is called the trajectory of the


point

in the Cartesian coordinate system

O
ei .
In

the

spatial

(4.142) at fixed

description,

Xk

the

Figure 1.14.

Trajectory of material
point

trajectory

is a solution of the kine-

matic equation (4.10):

dxi /dt = vi (xj , t),

< t 6 t0

(4.143)

58

Chapter 1. Kinematics of Continua

with the initial condition

Here

i ,
e

vi (xj , t)

xi = xi .

t=0:

are the velocity components with respect to the Cartesian basis

which are assumed to be known.

i be given. Fix a time t, and take


v(xj , t) = vi e
i
k
point M1 with Eulerian coordinates x1 and Lagrangian coordinates X1 . Then
streamline passing through the point M1 is the curve
Let a field of velocities

xi = xi (X k , ),
which at its every point

xi

1 6 6 2 ,

a
a

(4.144)

has a tangent being parallel to the velocity

v(xi , t)

at

the considered point and at the considered time. The equation of the streamline
has the form

dxi /d = vi (xj , t),

1 < 6 2 ,

(4.145)

xi = xi1 .

= 1 :

Thus, the trajectory of a material point and the streamline are described, in
general, by different equations, and so they are not coincident.
However, if the motion of a continuum is steady-state within time interval

t1 6 t 6 t0 ,

then in Eulerian description all the partial derivatives of all values,

describing the motion, with respect to time vanish, in particular

v(xi , t)/t = 0.

So the trajectory equations (4.143) and the streamline equation (4.145) become
coincident within the interval

M1 :

t1 6 t 6 t0 ,

dxi /d = vi (xj ),

if they have at least one common point

t 1 = 1 < 6 2 = t 0 ,

(4.146)

xi = xi1 = xi (X1k , t1 ).

= 1 :

In other words, in the steady-state motion a material


point

moves along a streamline: at time

t = t1 its
M1

coordinates are the same as coordinates of point


at parameter value

= 1 , and at time t = t2 they are


M2 at parameter

the same as coordinates of point


value

= 2

(Figure 1.15).

Multiplying the equation (4.145) by the basis


vectors

Figure 1.15. The streamline

i ,
e

we can rewrite the streamline equation in

the vector form

dx = v(x, t)d , 1 < 6 2 ,


x = x1 ,

= 1 .

(4.147)

Let us define now a vortex line passing through a point


curve, which at its every point
vector

(xj , t)

xi

M1 :

this is a

has a tangent being parallel to the vorticity

at the considered point and at the fixed time

t.

The vortex line is

described by the equation

dx = (x, t)d ,
x = x1 ,

1 < 6 2 ,
= 1 .

(4.148)

59

1.4. Rate Characteristics of Continuum Motion

1.4.13. Stream Tubes and Vortex Tubes. Consider a curve L in coordinates


xi and draw a streamline through each point of the curve L. If L is not a
streamline itself, then we get a surface

v ,

at each point of which the velocity

lies on the tangent plane to the surface. This surface is called the stream

surface.

Let

fv (xi ) = 0

(4.149)

be the equation of the stream surface. Since the vector


surface

[12], so it is orthogonal to the velocity

v,

is normal to the

i.e. we have the relation

v fv = 0,

(4.150)

which is a partial differential equation for determination of the function

i
the known velocity field v(x , t) at fixed t.
If a curve

f (xi )

by

is closed, then the set of streamlines drawn through its points

is called the stream tube.

L be
L, we
f (xi ) = 0.

Let a curve

not a vortex line. Drawing a vortex line through each point

of the curve

obtain the vortex surface

equation

This relation is a solution of the differential equation

which is described by the

f = 0.
If

is a closed curve, then the surface

(4.151)

is called the vortex tube.

Exercises for 1.4.


Exercise 1.4.1. Show that the tensors

QU

and

QV

are orthogonal.

Exercise 1.4.2.

Using formulae (4.104) and (2.57), show that for the coefficient

determined by formula (2.54) we have the following relationship:

k = k(n D n).
Exercise 1.4.3.

Using formulae (2.57), (4.99) and the result of Exercise 1.4.2, show

that a rate of changing the normal

is determined as follows:

n = n n L,
Exercise 1.4.4.

= n D n.

Using the results of Exercise 1.4.2, show that for the coefficient

determined by formula (2.54) we have the following equation:

k = k(n D n).
Exercise 1.4.5.

Show

that

the

transformations

(4.65)

of

infinitesimal

tension

compression (4.65) and infinitesimal rotation (4.66) are commutative up to terms of


order

(dt)2 :
Q AD dx = A QD dx,

while transformations of a small neighborhood determined by the tensors

and

are not commutative in general.

Exercise 1.4.6. Prove Theorem 1.14.

and

or

60

Chapter 1. Kinematics of Continua

Exercise 1.4.7. Using the representation (3.6) for tensors

U and V

and formulae (4.90)

and (4.92), show that rates of stretch tensors are expressed as follows:

=
U

3
X

p p + U U U U ,

=
V

=1

3
X

p p + V V V V .

=1

Exercise 1.4.8. Show that expressions for rates of the Hencky tensors (3.31) have the
form
3
X


p p + U H H U ,

H =

=1

3
X

H =

=1

p p + V H H V .

Exercise 1.4.9. Using representations (4.114) and (4.124) for tensors

and

V ,

and

also equation (4.120) and the result of Exercise 1.4.8, show that rates of the Hencky
tensors can be expressed in the form

H =

3
X

X
2

D p p +

,=1

,=1

lg

1 D p p ,

6=

=D+
H

X
2

,=1

2 2

lg

1 (p D p )p p .

6=

Exercise 1.4.10. Show that expressions for


rewritten as follows:

H = 4 XH D,

and

derived in Exercise 1.4.9 can be

= 4 XH D,
H

where the following fourth-order tensors are denoted:

XH

XH = XH ijkl pi pj pk pl , 4 XH = XH ijkl pi pj pk pl ,

2 lg
kl , 6= ,

kl = (1/2)(k l + l k ).
ijkl =

,
kl
4

Exercise 1.4.11.
and (4.124) for

Show that relations (4.114) and (4.122) for


and equation (4.129) for

U = 4 U D,

U ,

equations (4.123)

can be rewritten as follows:

V = 4 V D + W,

e D + W,
= 4

where
4

U = U ijkl pi pj pk pl ,

V = V ijkl pi pj pk pl ,

e =
e ijkl pi pj pk pl ,

2
+
kl , 6= ,
kl ,
=
V kl =
0,

= ,
4

U kl

0,

6= ,
= ,

61

1.5. Co-rotational Derivatives

e = V (l l l l ),

6= , 6= ,

2 2

if

(or

= ),

Exercise 1.4.12.

then the first (or the second) summand vanishes.


Using the definitions (4.45) and (4.46) of the tensors

and also the properties of unit tensors, show that


formulae

e L,
W=

D = L,

D, W

and

and

W,

are connected by the

e = (1/2)(III II ).

1.5. Co-rotational Derivatives


1.5.1. Definition of Co-rotational Derivatives.

da/dt

Besides the total derivative

introduced in paragraph 1.4.1 and partial derivative of vectors and tensors

with respect to time

a/t,

so-called co-rotational derivatives are of great

importance in continuum mechanics. They determine rates of changing tensors


relative to some moving basis

hi ,

i.e. the relative rates.

K(t) there be some moving bases hi or hi


i
i
and arbitrary varying scalar (X , t), vector a(X , t) and second-order tensor
i
T(X , t) fields with the following components with respect to the bases:
Let in a actual configuration

a = ai hi = ai hi ,
ij

T = T hi hj = Tij h h =
Since any scalar function

(X i , t)

T ij hi

(5.1)
j

j i

h = Ti h hj .

(5.2)

is not connected to any basis (moving or

fixed), it is evident that the co-rotational derivative of the function must be


coincident with the total derivative with respect to time:

h = .

(5.3)

a and a tensor T we can introduce co-rotational derivatives ah


h
and T as vectors or tensors, components of which with respect to the same
For a vector

basis

hi

ah =

dai
hi ,
dt

If we consider the basis

hi ,

rotational derivatives:

Th =

hi .

(or a tensor

T)

d ij
T hi hj .
dt

d
Tij hi hj .
dt

(5.6)

ah

(or

Th )

determines the rate of varying

for an observer moving together with the basis

For the observer, the basis

determine the rates of changing

hi .

(5.4)

(5.5)

hi

is fixed, and hence in (5.4) the basis is not

differentiated with respect to time. In a similar way, the derivatives


the basis

components,

dai i
h,
dt

aH =

Thus, the co-rotational derivative

and tensor

then for the basis we can determine other co-

TH =
a vector

coincide with rates of changing vector

respectively:

and

aH

and

TH

for an observer moving together with

62

Chapter 1. Kinematics of Continua

We can determine the co-rotational derivatives of a second-order tensor


mixed moving dyadic bases

Td =

hi hj

and

d i
T hi hj ,
dt j

Since vector components

ai

and

a =ah
and tensor components

i
of (5.2) by h or

ij

=h Th

hj ,
,

Tij , T ij , T ij

TD =

ai
i

hi hj ,

in

respectively:

d j i
T h hj .
dt i

(5.7)

can always be expressed in the form

and

ai = a hi ,
Ti j ,

(5.8)

with the help of the scalar product

can be written as follows:

Tij = hi T hj ,

T ij = hi T hj ,

Ti j = hi T hj ,
(5.9)

so rates of changing vector and tensor components in (5.4), (5.5) and (5.7) can
be represented in the explicit form:

dai
da
dhi
=
hi + a
,
dt
dt
dt
and also

dai
da
dh
=
hi + a i ,
dt
dt
dt

dT ij
dT
dhi
dhj
= hi
hj +
T hj + hi T
,
dt
dt
dt
dt
dTij
dT
dh
dh
= hi
hj + i T hj + hi T j ,
dt
dt
dt
dt
i
d i
d
dh
dh
T = hi T hj +
T hj + hi T j ,
dt j
dt
dt
dt
d
dh
d
d j
T = hi T hj + i T hj + hi T hj .
dt i
dt
dt
dt

Here total derivatives

da/dt

and

dT/dt

(5.10)

(5.11)

are determined by the rules (4.7) and

dhi /dt and dhi /dt are


j
defined by the choice of basis hi or h .
j
Taking different bases as hi and h , we get different co-rotational derivatives.

(4.12), respectively. Rates of changing basis vectors

Let us consider the most widely used bases.

1.5.2. The Oldroyd Derivative(hi = ri ). If we choose the general local vector


h
Ol (or Th = TOl ) determines the rate
basis ri as hi , then the derivative a = a
i
of changing a (or T) relative to the Lagrangian coordinate system X moving
together with the continuum. This derivative is called the Oldroyd derivative.
The derivative

dri /dt

is determined as follows:

dhi
dr
2x
v
v
= i =
=
= ri rj
= ri v = ( v)T ri . (5.12)
dt
dt
tX i
X i
X j

hi we consider the reciprocal local basis


i
derivative of which dr /dt with respect to time has the form
In this case, as a basis

d
d i
(r ri ) = E =
dt
dt
or

0,

dr
dri
ri = ri i = ri ( v)T ri .
dt
dt

ri ,

the

(5.13)
(5.14)

63

1.5. Co-rotational Derivatives

Multiplying the equation by


j

rj

from the right, we get

dh
dr
=
= rj ( v)T = ( v) rj .
dt
dt
On substituting the expressions (5.15) for the derivatives

(5.15)

dhi /dt

and

into (5.10), we get the formula for the Oldroyd derivative in the basis

aOl =
TOl =

ri :

dai
da i
d
ri =
r ri a ( v) ri ri = a a v,
dt
dt
dt

dhj /dt
(5.16)

dTij
dT ij
d
ri rj = ri
rj = ri ri T rj rj
dt
dt
dt

ri ri ( v)T T rj rj ri ri T ( v) rj rj =
=

d
T T v ( v)T T.
dt

Here we have taken into account that

ri ri = E.

(5.17)

Thus, we have proved the

following theorem.

Theorem 1.29.

The Oldroyd derivative is related to the total derivative with

respect to time as follows (for a vector

and for a tensor

T,

respectively):

T v ( v)T T.
TOl = T

aOl = a a v,

(5.18)

1.5.3. The CotterRivlin Derivative (hi = ri ). If we choose the reciprocal


i
i
H (or TH ) characterizes
local basis r as a moving basis h , then the derivative a
i
the rate of changing a (or T) relative to the basis r moving together with
i
the Lagrangian coordinate system X . This derivative is called the Cotter-Rivlin
derivative.

Because of formulae (5.10) and (5.15), we get the following theorem.

Theorem 1.30.

The CotterRivlin derivative is related to the total derivative

as follows (for a vector

and for a tensor

aH aCR =
TH TCR =

T,

respectively):

dai i
r = a + ( v) a,
dt

(5.19)

dTij i
+ v T + T ( v)T .
r rj = T
dt

1.5.4. Mixed Co-rotational Derivatives. Since any vector

(5.20)

is defined by its

hi or
hj , so for the vector in the moving bases we can determine only two co-rotational
components with respect to a vector basis, for example, in a moving basis
derivatives: by Oldroyd and by CotterRivlin.
Any second-order tensor

is defined by its components with respect to a

dyadic basis. Therefore, besides the Oldroyd and CotterRivlin derivatives, which
specify the rates of changes of a tensor

ri rj ,

in moving dyadic bases

ri rj

and

by formulae (5.7) we can determine two more derivatives in moving

mixed dyadic bases:

Td =

dT ij
ri rj ,
dt

TD =

d j i
T r rj .
dt i

(5.21)

64

Chapter 1. Kinematics of Continua

On substituting the expressions (5.12) and (5.15) into (5.11), we get the
following formulae for the rates of changing mixed components of the tensor

T:

d i
rj ri ( v)T T rj + ri T ( v)T rj ,
T = ri T
dt j
d j
rj + ri ( v) T rj ri T ( v) rj .
T = ri T
dt i

(5.22)

Having substituted (5.22) into (5.21), we get the following theorem.

Theorem 1.31.

The mixed derivatives (5.21) are connected to the total deriva-

tive by the relations

L T + T L,
Td = T

+ LT T T LT .
TD = T

(5.23)

The derivatives (5.21) are called the left and right mixed co-rotational
derivatives, where

L = ( v)T

is the velocity gradient (see (4.43)).

It should be noticed that, unlike other co-rotational derivatives considered in


this paragraph, the mixed derivatives

Td

when they are applied to a symmetric

TD do not form a symmetric tensor


tensor T. This fact explains a scarcer

and

application of mixed derivatives in continuum mechanics.

1.5.5. The Derivative Relative to the Eigenbasis pi of the Right Stretch

Tensor. If we choose the eigenbasis pi of the right stretch tensor U as

i
a moving basis hi , then, since pi are orthonormal, we get that h and hj

i
are coincident: h = h , = 1, 2, 3, and |h | = 1. At every time, the moving

coordinate system defined by the trihedron pi executes an instantaneous rotation,


which is characterized by the spin U (4.89), and due to (4.90) we have

dhi
= pi = U pi = pi T
U = pi U .
dt

(5.24)

On substituting (5.24) into (5.11), we get

ah aU =
Th TU =

dai
pi = a pi pi + a U pi pi = a + a U ,
dt

T ij
dT ij
p
pi pj = pi
pj = pi pi T
pj
dt
dt

pi pi U T pj pj + pi pi T U pj pj =
U T + T U .
=T
The co-rotational derivative of a vector

(or a tensor

(5.25)

T)

(5.26)

determined by

(5.26) is called the right derivative relative to the eigenbasis.


Thus, we have proved the following theorem.

Theorem 1.32.

The right derivative relative to the eigenbasis is connected to

the total derivative as follows (for a vector

aU = a + a U ,

and for a tensor

T,

U T + T U .
TU = T

respectively):

(5.27)

65

1.5. Co-rotational Derivatives

1.5.6. The Derivative in the Eigenbasis (hi = pi ) of the Left Stretch


Tensor. Take the eigenbasis pi of the left stretch tensor V as a moving basis

hi

and define the following co-rotational derivatives

aH aV =

dai
pi ,
dt

dT ij
pi pj ,
dt

TH TV =

(5.28)

called the left derivatives in the eigenbasis.

Theorem 1.33.

The left derivatives (5.28) in the eigenbasis are connected to

the total derivative with respect to time by the following relations (for a vector

and for a tensor

T,

respectively):

V T + T V .
TV = T

a = a V a,
H

(5.29)

A proof follows from (5.5), (5.10) and (5.11), because from (4.92) we have

dhi
= p i = V pi .
dt
Since the bases

pi

and

pi

are orthonormal, all the co-rotational derivatives

relative to the mixed dyadic bases


respectively.

(5.30)

pi pi , pi pi

coincide with

TU

or

TV ,

1.5.7. The Jaumann Derivative (hi

the

deformation rate tensor as a moving basis

the

basis

qi

is also orthonormal and

= qi ). If we choose the eigenbasis of


hi = qi (it should be noted that
i
coincides with q ), then from (5.4) we get

the

co-rotational Jaumann derivatives:

ah aJ =
Theorem 1.34.

dai
qi ,
dt

Th TJ =

dT ij
qi qj .
dt

The Jaumann derivatives (5.31) are connected to the total

derivatives with respect to time by the relations (for a vector

T,

(5.31)

respectively)

and for a tensor

aJ = a + a W,
W T + T W.
TJ = T

(5.32)
(5.33)

According to the relationship (4.94), we get

dhi
= q i = W qi ,
dt
therefore, due to formulae (5.5) and (5.10) we find

aJ = a qi qi + a W qi qi = a + a W.
In a similar way, we can prove the relation (5.33).

1.5.8. Co-rotational Derivatives in a Moving Orthonormal Basis.

Let

hi

be a moving orthonormal basis. In this case we denote co-rotational derivatives


by the following way:
basis

hi ,

ah aQ

the total derivatives of

Th TQ . Due to orthonormalization of the


a and T with taking account of (5.1), (5.2) and

and

(5.4) can be written as follows:

da
dai
dh
=
hi + ai i = aQ + ai h hi ,
dt
dt
dt

(5.34)

66

Chapter 1. Kinematics of Continua

ij
= dT hi hj + T ij dhi hj + hi dhj =
T
dt

dt

dt

= TQ + T ij h hi hj hi hj h .
i of the moving basis, where
Here we have used formula (4.86) for derivative h
h is the vorticity vector giving a rotation of the basis hi relative to the fixed
i (see (4.78) and (4.85)). With taking account of (5.1) and (5.2), formulae
basis e
(5.34) can be written in the form

h T + T h.
TQ = T

aQ = a h a,
It should be noticed that if

a = h,

(5.35)

then

h = hh ,
because

h h = 0

(5.35a)

due to properties of the vector product.

i
h

1.5.9. Spin Derivative. Take an arbitrary orthonormal basis


of a continuum in

the basis

i
h

K.

at a point

The trihedron must have the only property that at any time

rotates with the instantaneous angular rate, which is equal to the

rotation rate of the trihedron

pi

relative to the trihedron

pi .

As shown in 1.4.10,

the instantaneous rotation of the trihedron is characterized by the spin tensor

OT
=O

determined by (4.97).

Then we can define the co-rotational derivative in the basis, which is called
the spin derivative (of a vector

and of a tensor

Theorem 1.35.

The

spin

derivative

(5.36)

dT ij
j.
hi h
dt

(5.37)

is

respect to time as follows (for a vector

respectively):

da
hi ,
dt

ah aS =
Th TS =

T,

related

to

the

total

and for a tensor

T,

derivative

aS = a + a ,
T + T .
TS = T
H

with

respectively):

(5.38)
(5.39)

A proof of Theorem 1.35 follows from (5.10), (5.11) and the relation

i
h
i,
=h

(5.40)

which is a consequence of (4.84). The relation (5.39) follows from (5.11).

1.5.10. Universal Form of the Co-rotational Derivatives.

On comparing

formulae (5.18), (5.19), (5.20), (5.27), (5.29), (5.33), (5.38) and (5.39), we can
notice that all the representations of the co-rotational derivatives and also the
total derivative with respect to time can be written in the universal form:

Zh T T ZT ,
ah = a Zh T,
Th = T
h
h = { , Ol, CR, U , V , J , S },
where tensors

Zh

have the following representation for different

Zh = { 0, L, L

U , V , W, },

(5.41)
h:
(5.42)

67

1.5. Co-rotational Derivatives

h = { , Ol, CR, U , V , J , S }.
U , V

Since tensors

and

are linearly expressed in terms of W and D


Zh can be written as linear functions of W and

(see Exercise 1.4.11), so tensors

D:

Zh = 4 ZDh D + 4 ZW h W.
4

Table 1.1 gives expressions for fourth-order tensors


tensors

U , V
4

and

Ol

CR

ZDh

III

III

ZW h

III

III

Eh

ZDh

ZW h ,

and

Eh

for

where

are defined in Exercise 1.4.11.

T a b l e 1.1. Expressions of tensors


different co-rotational derivatives

(5.43)

ZDh , 4 ZW h

U
4

and

III

III

III

S
4

1.5.11. Relations between Co-rotational Derivatives of Deformation Rate


Tensors and Velocity Gradient. In paragraph 1.4.11 we have derived the
relationships between rates of deformation tensors and velocity gradient

L.

Similar connections also exist between co-rotational derivatives of the tensors


and

L.

Let us establish them.

On substituting representations (4.103), (4.104) and (4.130) for rates

g , (g1 ) ,

1 ) into formula (5.41), we get


and (V

, J ,
A

Ah = D (Zh + LT ) A A (ZT
h + L),
T
Jh = D (Zh L) J J (ZT
h L ),

gh = (Zh + LT ) g g (ZT
h + L),
(g

1 h

) = (Zh L) g

(ZT
h

(5.44)
T

L ),

T
T
Vh = (Zh (L + )) V V (ZT
h (L + )),
1

T
(V1 )h = (Zh ( LT )) V1 V1 (ZT
h ( L)),

h = { , Ol, CR, U , V , J , S }.
From these relationships we can find the following expressions:

(V E)h = D 4 Eh D (Zh + ZT
h )
1

T
T
(Zh ( + L)) (V E) (V E) (ZT
h ( + L )),

(5.45)

68

Chapter 1. Kinematics of Continua

(E V1 )h = D + 4 Eh D + Zh + ZT
h
1

T
(Zh ( LT )) (E V1 ) (E V1 ) (ZT
h ( L)).
Here we have denoted the co-rotational derivative of the metric tensor by

Eh = 4 Eh D.
The tensor

Eh

(5.46)

differs from zero-tensor only when

h = {CR, Ol}

(see Exercise

1.5.3), its expressions are given in Table 1.1 (see paragraph 1.5.10).
Since tensors

Zh

and

are linearly expressed in terms of

and

(see formulae (5.43), (4.129), (4.122)(4.124)), so on the right-hand sides of


equations (5.44) there are also linear functions of

and

D,

their explicit

expressions will be given in 3.2.22.

Exercises for 1.5.


Exercise 1.5.1. Show that the mixed co-rotational derivatives, the left and the right corotational derivatives relative to the eigenbasis and also the Jaumann and spin derivatives
satisfy the differentiation rules of scalar products:

(A B)h = Ah B + A Bh ,
(A)h = h A + Ah ,

(a A)h = ah A + a Ah ,
h = {, d, D, U , V , J , S},

and the Oldroyd and Cotter-Rivlin derivatives do not satisfy this rule.

Exercise 1.5.2.

Show that for the co-rotational derivatives, the following rules of


differentiation of scalar products of two vectors a and b and also of two tensors T and
B remain valid:

(a b)h = (a b) = ah b + a bh ,

h = {U , V , J , S};

(T B)h = (T B) = Th B + T Bh ,

h = {d, D, U , V , J , S}.

Exercise 1.5.3.

Show that the following co-rotational derivatives of the unit tensor


give the zero-tensor:

Eh = 0,

and the Oldroyd and CotterRivlin

h = {, d, D, U , V , J , S},
derivatives of E are different

ECR = 2D,

from zero:

EOl = 2D.

Exercise 1.5.4.

Using formulae (4.103) and (5.20), show that the CotterRivlin


derivatives of the left Almansi deformation tensor A and of the left Almansi measure g
have the form
ACR = D,
gCR = 0.

Exercise 1.5.5.

Using formulae (4.103) and (5.18), show that the Oldroyd derivatives

of the right CauchyGreen tensor


form

Ol

Exercise 1.5.6.
also

and

eigenbasis

G1

and of the right deformation measure

= D,

(g

Using the expressions for the

g 1

have the

1 Ol

= 0.
tensors U

(3.6),

and

(3.29), and

(3.29), show that we can write the right derivative relative to the

in the form

UU =

3
X

=1

p p ,

CU =

3
X

=1

p p ,

69

1.5. Co-rotational Derivatives

U =

3
X


p p ,

=1

3
X


p p ,

=1

(G1 )U = 2U .

GU = 2CU ,
Exercise 1.5.7.

(U1 )U =

Using the expressions for the tensors

(3.6), for

and

(3.29),

and also for g and g


(3.29), show that we can write the left derivative relative to the
eigenbasis p in the form

VV =

3
X

p p ,

AV =

=1

JV =

3
X

3
X

=1

p p ,

gV = 2AV ,

p p ,

(g1 )V = 2JV .

=1

Exercise 1.5.8.
tensor

Show that the Oldroyd and Jaumann derivatives of a second-order


are connected by the relationship

TOl = TJ T D D T.
Exercise 1.5.9.

Show that if for an arbitrary symmetric tensor


derivatives are equal to zero:

Th = 0,

its co-rotational

h = {d, D, U , V , J , S},

then the first invariant of the tensor:

I1 (T) = T E

has its stationary value, i.e.

I1 (T) = 0.
Show that for the co-rotational Oldroyd and CotterRivlin derivatives this statement is
not valid.

Exercise 1.5.10.

Using the results of Exercise 1.4.3, show that the co-rotational


derivatives of the normal vector n satisfy the following relations:

nCR = n,

= n D n,

nOl = n n L L n,

nJ = n n D.

Exercise 1.5.11.

Show that the following co-rotational derivatives of a symmetric


tensor give a symmetric tensor:
if

A = AT ,

then

(Ah )T = Ah , h = {U , V , J , S ,

Ol, CR},

and also a skew-symmetric tensor, if they are applied to a skew-symmetric tensor:


if

B = BT ,

then

(Bh )T = Bh , h = {U , V , J , S , Ol, CR}.

The mixed co-rotational derivatives

h = d, D

have no such properties.

http://www.springer.com/978-94-007-0033-8

You might also like