Вы находитесь на странице: 1из 182

CH.2.

DEFORMATION AND
STRAIN
Multimedia Course on Continuum Mechanics
Overview
 Introduction Lecture 1
 Deformation Gradient Tensor
 Material Deformation Gradient Tensor Lecture 2
 Inverse (Spatial) Deformation Gradient Tensor
Lecture 3

 Displacements Lecture 4
 Displacement Gradient Tensors
 Strain Tensors
 Green-Lagrange or Material Strain Tensor Lecture 5
 Euler-Almansi or Spatial Strain Tensor
 Variation of Distances
 Stretch Lecture 6
 Unit elongation
 Variation of Angles Lecture 7

2
Overview (cont’d)
 Physical interpretation of the Strain Tensors Lecture 8
 Material Strain Tensor, E
 Spatial Strain Tensor, e Lecture 9
 Polar Decomposition Lecture 10
 Volume Variation Lecture 11
 Area Variation Lecture 12
 Volumetric Strain Lecture 13
 Infinitesimal Strain
 Infinitesimal Strain Theory
 Strain Tensors
 Stretch and Unit Elongation Lecture 14
 Physical Interpretation of Infinitesimal Strains
 Engineering Strains
 Variation of Angles

3
Overview (cont’d)
 Infinitesimal Strain (cont’d)
 Polar Decomposition Lecture 15
 Volumetric Strain
 Strain Rate
 Spatial Velocity Gradient Tensor Lecture 16
 Strain Rate Tensor and Rotation Rate Tensor or Spin Tensor
 Physical Interpretation of the Tensors
 Material Derivatives Lecture 17
 Other Coordinate Systems
 Cylindrical Coordinates Lecture 18
 Spherical Coordinates

4
2.1 Introduction
Ch.2. Deformation and Strain

5
Deformation
 Deformation: transformation of a body from
a reference configuration to a current configuration.

 Focus on the relative movement of a given particle w.r.t. the particles


in its neighbourhood (at differential level).

 It includes changes of size and shape.

6
2.2 Deformation Gradient Tensors
Ch.2. Deformation and Strain

7
Continuous Medium in Movement
Ω0: non-deformed (or reference) Ω or Ωt: deformed (or present)
configuration, at reference time t0. configuration, at present time t.
X : Position vector of a particle at x : Position vector of the same particle
reference time. at present time.

ϕ ( X,t )

t0 Q t
Reference or Ωt
dX Q’
non-deformed dx

Ω0 P P’
X x Present or
deformed

8
Fundamental Equation of Deformation
 The Equations of Motion:
not
xi i ( X1, X 2 , X 3 , t )
ϕ= xi ( X 1 , X 2 , X 3 , t ) i ∈ {1, 2,3}
not
= ( X, t ) x ( X, t )
x ϕ=

 Differentiating w.r.t. X :
 ∂xi ( X, t ) not
 dxi = dX j Fij ( X, t ) dX j i, j ∈ {1, 2,3}
 ∂X j

 ∂x ( X, t ) not Fundamental equation
 d=
x X F ( X, t ) ⋅ dX
⋅ d= of deformation
∂X

(material) deformation
gradient tensor

9
Material Deformation Gradient Tensor

 The (material) deformation gradient tensor: REMARK  ∂ 


 
The material Nabla operator  ∂X 1 
 not  ∂ 
F (=
X, t ) x( X, t ) ⊗ ∇ is defined as: ∇ ≡ ∂ eˆ i ∇  = 
∂X i  ∂X 2 
 not ∂x  ∂ 
=  Fij
i
i, j ∈ {1, 2,3}  
 ∂X 3 
 ∂X j
 ∂x1 ∂x1 ∂x1 
 
 x1   ∂X 1 ∂X 2 ∂X 3 
 ∂ ∂ ∂   ∂x2 ∂x2 ∂x2 
[F ] = x ⊗ ∇  = x2  

 = 
∂X ∂X 2 ∂X 3   ∂X 1 ∂X 2 ∂X 3 
 x3   1
 ∂x3 ∂x3 ∂x3 
=  x  =∇ T  
 ∂X 1 ∂X 2 ∂X 3 

 F(X,t):
 is a primary measure of deformation
 characterizes the variation of relative placements in the neighbourhood of a
material point (particle).
=dx F ( X, t ) ⋅ dX
10
Inverse (spatial) Deformation Gradient
Tensor
 The inverse Equations of Motion:
not
( x1 , x2 , x3 , t ) X i ( x1 , x2 , x3 , t )
X i ϕ= i
−1
i ∈ {1, 2,3}
not
= ( x, t ) X ( x, t )
X ϕ= −1

 Differentiating w.r.t. x :
∂X i ( X, t ) not
=dX i = dx j Fij−1 ( x, t ) dx j i, j ∈ {1, 2,3}
∂x j
∂X ( x, t ) not
=
dX =⋅ dx F −1 ( x, t ) ⋅ dx
∂x

Inverse (spatial) deformation


gradient tensor

11
Inverse (spatial) Deformation Gradient
Tensor
 The spatial (or inverse) deformation gradient tensor:
= dX F −1 ( x, t ) ⋅ dx

∂
 F ( x, t ) ≡ X(x, t ) ⊗ ∇
−1 REMARK x 
 The spatial Nabla  1
 −1 ∂X i ∂
=
 Fij i, j ∈ {1, 2,3} operator is defined as: [∇] = 
 ∂x j ∂  x2 
∇≡ eˆ i ∂
 ∂X 1 ∂X 1 ∂X 1  ∂xi
   
= ∇T  ∂x1 ∂x2 ∂x3   x3 
 X1 
  ∂ ∂ ∂   ∂X 2 ∂X 2 ∂X 2 
F  = [ X ⊗ ∇ ] =  X 2   ∂x
−1
= 
∂x2 ∂x3   ∂x1 ∂x2 ∂x3 
 X 3   1
 ∂X 3 ∂X 3 ∂X 3 
 F-1(x,t): =  X   
 ∂x1 ∂x2 ∂x3 
 is a primary measure of deformation
 characterizes the variation of relative placements in the neighbourhood of a
spatial point.
 It is not the spatial description of the material deformation gradient tensor

12
Properties of the Deformation
Gradients
 The spatial deformation gradient tensor is the inverse of the material
deformation gradient tensor:

∂xi ∂X k ∂xi
= = δ ij F ⋅ F −1 = F −1 ⋅ F = 1
∂X k ∂x j ∂x j

 If F is not dependent on the space coordinates, F( X, t ) ≡ F ( t ) the


deformation is said to be homogeneous.
 Every part of the solid body deforms as the whole does.
 The associated motion is called affine.

∂x
 If there is no motion, =
x X and =
F −1
= F= 1 .
∂X

13
Example
Compute the deformation gradient and inverse deformation gradient tensors
for a motion equation with Cartesian components given by,
 X + Y 2t 
[ x] Y (1 + t ) 
=
 Zet 
 

1.
Using the results obtained, check that F ⋅ F −1 =

14
 X + Y 2t 
 
[ ]  ( )
Example - Solution =x Y 1 +
 Zet 

t

The Cartesian components of the deformation gradient tensor are,


 X + Y 2t  1 2Yt 0 
  ∂ ∂ ∂   
F ( X, t )  =x ⊗ ∇ ≡ [ x ] ∇  =Y (1 + t )  
T
, , = 0 1 + t 0
 ∂X ∂Y ∂Z   
 Ze  t
0 0 e  t
 

The Cartesian components of the inverse motion equation will be given by,

 y 2t 
 X= x −   2 yt 
( + )
2
 1 t   1 0
  (1 + t )
 y     
[ X=] ϕ −1 ( x, t =
)  =
   f (=
( ) x, t )  0 1 + t
Y  F X ( x , =
t ), t 0
 1+ t 
  0 et 
 Z = ze − t   f ( x ,t )  
0

 
   

15
Example - Solution
The Cartesian components of the inverse deformation gradient tensor are,
 2 yt 
 1 − 0
( + )
2
 1 t 
 1 
F −1 ( x, t )  = 0 0
 1+ t 
 
 
0 0 e − t 

1:
And it is verified that F ⋅ F −1 =
 2 yt   2 yt 2 yt 
  1 − 0 1 − + 0 
0  
2 yt
(1+ t) (1 + t ) (1 + t )
2 2 2
 1
(1 + t )     1 0 0
   1   1+ t  0 1 0= 
F ⋅ F −=
1
0 1 + t 0  ⋅ 0 0=  0 0 =    [1]
0 1+ t 1+ t
0 et      0 0 1 
     
  0  0 
 0 e−t   0 et e − t 

16
2.3 Displacements
Ch.2. Deformation and Strain

17
Displacements
 Displacement: relative position of a particle, in its current (deformed)
configuration at time t, with respect to its position in the initial
(undeformed) configuration.
 Displacement field: displacement of all the particles in the continuous
medium.
 Material description (Lagrangian form):
( X, t ) x ( X, t ) − X
U= t0 t

U i ( X, t ) = xi ( X, t ) − X i i ∈ {1, 2,3} Ω
P U=u P’
 Spatial description (Eulerian form): Ω0 X
x
u ( x, t )= x − X ( x, t )
ui ( x, t ) =
xi − X i ( x, t ) i ∈ {1, 2,3}

18
Displacement Gradient Tensor
U= ( X, t ) x ( X, t ) − X  Taking partial derivatives of U w.r.t. X :
 ∂U i ( X, t ) ∂xi ( X, t ) ∂X i
U i ( X , t ) = xi ( X , t ) − X i i ∈ {1, 2,3}
def
= − = Fij − δ ij = J ij
∂X j ∂X j ∂X

  j
 ∂U i δij
 J = = Fij − δ ij i, j ∈ {1, 2,3} Fij
∂X j
ij

 Material Displacement
 def
Gradient Tensor
 (
J X , t ) = U ( X, t ) ⊗ ∇ = F − 1
 Taking partial derivatives of u w.r.t. x :
u ( x, t )= x − X ( x, t ) ∂ui (x, t ) ∂xi ∂X i (x, t ) def
 = − = δ ij − Fij − δ ij = jij
−1

ui ( x, t ) =
xi − X i ( x, t ) i ∈ {1, 2,3} ∂x j ∂x j
 
∂x j
 
δij Fij−1
 ∂ui
 ij ∂x =
j = δ ij − Fij−1 i, j ∈ {1, 2,3} Spatial Displacement Gradient
 j
Tensor
 def
 j ( x, t ) = u ( x, t ) ⊗ ∇ =1 − F
−1

REMARK If motion is a pure shifting: ∂x


x( X, t ) =X + U(t ) ⇒ F= = 1=F −1 and j =J=0
. ∂X
19
2.4 Strain Tensors
Ch.2. Deformation and Strain

20
Strain Tensors
 F characterizes changes of relative placements during motion but is not
a suitable measure of deformation for engineering purposes:
 It is not null when no changes of distances and angles take place, e.g.,
in rigid-body motions.

 Strain is a normalized measure of deformation which characterizes the


changes of distances and angles between particles.
 It reduces to zero when there is no change of distances and angles
between particles.

21
Strain Tensors
Consider
F ( X,t )

t0 Q t d=
x F ⋅ dX
Ω dxi = Fij dX j
dS dX dx Q’
ds
d=
X F -1 ⋅ dx
Ω0 P P’
X x dX i = Fij−1dx j

 where dS is the length of segment dX : =dS dX ⋅ dX


 and ds is the length of segment dx : =
ds dx ⋅ dx

22
Strain Tensors
d=
X F -1 ⋅ dx d=
x F ⋅ dX
dX i = Fij−1dx j dxi = Fij dX j

=
dS dX ⋅ dX =
ds dx ⋅ dx

 One can write:


( ) = ⋅ = [ ] ⋅ [ ] = [ ⋅ ] ⋅ [ F ⋅ dX ] = dX ⋅ FT ⋅ F ⋅ dX
2 T T
 ds dx dx dx dx F dX

( ds ) dx
= = = =
2 T
k dx k F ki dX F
i kj dX j dX F F
i ki kj dX j dX i ik Fkj dX j
F
REMARK
 not
( dS ) = dX ⋅ dX =[ dX ] ⋅ [ dX ] = F ⋅ dx  ⋅ F ⋅ dx  = dx ⋅ F ⋅ F ⋅ dx
2 T −1 T −1 −T −1 The convention
not

T
(•) −1  =
(•) −T
( dS )2 dX
= = k dX k Fki−1 dxi=
Fkj−1 dx j dxi Fki−=
1 −1
Fkj dx j dxi Fik−T Fkj−1dx j is used.

23
Green-Lagrange Strain Tensor
( ds ) = dX ⋅ FT ⋅ F ⋅ dX ( dS =
) dX ⋅ dX
2 2

 Subtracting:
( ds ) − ( dS ) = dX ⋅ FT ⋅ F ⋅ dX − dX ⋅ dX = dX ⋅ FT ⋅ F ⋅ dX − dX ⋅ 1 ⋅ dX = dX ⋅ ( FT ⋅ F − 1 ) ⋅ dX = 2 dX ⋅ E ⋅ dX
2 2


def
= 2E

 The Green-Lagrange or Material Strain Tensor is defined:



) ( F ⋅ F − 1)
1 T
 (
E X,=t
2

 E ( X, t ) = 1 ( F F − δ ) i, j ∈ {1, 2,3}
 ij 2
ki kj ij

( )
 E is symmetrical:
(2
1 T
) 2 F ⋅ ( F ) − 1= 2 ( F ⋅ F − 1=) E
1 T T T T 1 T
T
E=
T
F ⋅ F − 1=

=Eij E ji i, j ∈ {1,2,3}

24
Euler-Almansi Strain Tensor
( ds )= dx ⋅ dx ( dS ) =dx ⋅ F −T ⋅ F −1 ⋅ dx
2 2

 Subtracting:
( ds ) − ( dS ) = dx ⋅ dx − dx ⋅ F −T ⋅ F −1 ⋅ dx = dx ⋅ 1 ⋅ dx − dx ⋅ F −T ⋅ F −1 ⋅ dx =
2 2
def
= 2e
= dx ⋅ (1 − F ⋅ F ) ⋅ dx = 2 dx ⋅ e ⋅ dx
−T −1

  
def
= 2e
 The Euler-Almansi or Spatial Strain Tensor is defined:

 e ( x, t ) =
1
2
( 1 − F −T
⋅ F −1
)

 e ( x, t ) =
 ij
1
2
( δ ij − Fki−1 Fkj −1 ) i, j ∈ {1, 2,3}

 e is symmetrical: e =
T 1
2
( −T −1 T 1 T
2
−1 T
( −T T 1
2
)
1 − F ⋅ F ) = 1 − ( F ) ⋅ ( F ) = (1 − F −T ⋅ F −1 ) = e

= eij e ji i, j ∈ {1, 2,3}


25
Particularities of the Strain Tensors
 The Green-Lagrange and the Euler-Almansi Strain Tensors are different
tensors.
 They are not the material and spatial descriptions of a same strain tensor.
 They are affected by different vectors (dX and dx) when measuring distances:

( ds ) − ( dS )= 2 dX ⋅ E ⋅ dX
= 2 dx ⋅ e ⋅ d x
2 2

 The Green-Lagrange Strain Tensor is inherently obtained in material


description, E = E ( X,t ) .
 =
By substitution of the inverse Equations of ( X ( x, t ) , t ) E ( x, t ) .
Motion, E E=

 The Euler-Almansi Strain Tensor is inherently obtained in spatial description,


e = e ( X,t ) .
 =
By substitution of the Equations of ( x ( X, t ) , t ) e ( X, t ) .
Motion, e e=

26
Strain Tensors in terms of Displacements
 Substituting F −1= 1 - j and F= J + 1 into
E=
2
( F ⋅ F − 1) and e= 1 (1 − F −T ⋅ F −1 ) :
1 T
2

 1 1
 E= (1 + J T ) ⋅ (1 + J ) − 1=  J + J T + J T ⋅ J 
 2 2

 Eij= 1  ∂U i ∂U j ∂U k ∂U k 
 + +  i, j ∈ {1, 2,3}
 2  ∂X j ∂X i ∂X i ∂X j 

 1
1 − (1 − jT ) ⋅ (1 − j) =  j + jT − jT ⋅ j
1
 e=  2 
 2
  ∂u 
 eij= 1  ∂ui + j − ∂uk ∂uk  i, j ∈ {1, 2,3}
 2  ∂x j ∂xi ∂xi ∂x j 

27
Example
For the movement in the previous example, obtain the strain tensors in the
material and spatial description.
 X + Y 2t 
[ x] Y (1 + t ) 
=
 Zet 
 

29
Example - Solution
The deformation gradient tensor and its inverse tensor have already been
obtained:  2 yt 
 1 − 0 
( + )
2
 1 t 
1 2Yt 0 
0 1 + t 0   1 
F=
  F −1 =
0 0
 1+ t 
0 0 e t
 
 
0 0 e − t 

The material strain tensor :


 1 0 0  1 2Yt 0   1 − 1 2Yt 0 
1  1 
=
E
2
(
1 T
F ⋅F −1 =) 2  
2Yt 1 + t 0  ⋅  0 1 + t 0  − 1=
    2  2Yt ( 2Yt ) + (1 + t ) − 1
2 2
0 = 
  0 e e − 1
  0 0 et  0 0 et  t t
 0
 
 0 2Yt 0 
1 
( ) + ( + ) −
2 2
 2Yt 2Yt 1 t 1 0 
2
0 0 e − 1
2t
 

30
Example - Solution
The spatial strain tensor :
 2 yt 
  2 yt   1−1 − 0 
1 − (1 + t )
2
  0 
 0  (1 + t )  
2
 
1 0
   
   1 2 
2
1    1  2 yt =
e=
1
(1 − F−T ⋅ F −1 )=  2 yt
1 − −
1
0  ⋅ 0

1
0  = −
 (1 + t ) 2
1− −
2 yt
 + 
  (1 + t ) 2   1 + t  
0

(1+ t) 1+ t 1+ t
2
2 2  2
        
  e − t  0 0 e−t    −t −t 

0 0
   0 0 1− e e
 
    
 2 yt 
 0 − 0 
(1 + t )
2
 
 2
 2 yt   1 2 
1  2 yt
= − 1−   − 0 
2  (1 + t ) 2  (1 + t ) 2   1 + t  
   
 0 0 1 − e −2 t 
 
 

31
 X + Y 2t 
[ x] Y (1 + t ) 
Example - Solution =
 Zet 
 

In conclusion, the material strain tensor is: spatial description


 2 yt 
material description  0 (1 + t )
0 
 0 2Yt 0   
1    2 yt 
2 
E ( X, t ) ( 2Yt ) + (1 + t ) − 1 0  1  2 yt
E ( x, t )  + (1 + t ) − 1 0 
2 2
=
2
 2Yt 
2 2  (1 + t )  (1 + t )  
0 0 e 2t − 1 Y=
y
 
  (1 + t )  0 0 e 2t − 1
 
And the spatial strain tensor is:  

 2 yt   
 0 − 0  −
2Yt
(1 + t )
2
 0 (1 + t )
0 
   
 2   
1  2 yt  2 yt   1 2   2Yt   1 
2 2
e ( x, t ) =
1  2Yt
 − 1−   −  0  e ( X, t ) = − 1 −   −   0 
2 (1 + t )  
 (1 + t )   1 + t  2  (1 + t )  (1 + t )   1 + t  
2 2
  y Y (1 + t )
=  
 0 0 1− e −2 t 
 0 0 1 − e −2t 
   
   
spatial description material description
Observe that E ( x, t ) ≠ e ( x, t ) and E ( X, t ) ≠ e ( X, t ).
32
2.5 Variations of Distances
Ch.2. Deformation and Strain

33
Stretch
 The stretch ratio or stretch is defined as:
def P´Q´ ds
stretch = λ T= λ=
t = (0 < λ < ∞)
PQ dS

t0 Q t t

dS dX Ω Q’ REMARK
dx
ds The sub-indexes (●)T and
(●)t are often dropped. But
Ω0 P P’
X x one must bear in mind that
stretch and unit elongation
always have a particular
direction associated to them.

34
Unit Elongation
 The extension or unit elongation is defined as:
def ∆ PQ ds − dS
unit elongation = ε=
T ε=
t =
PQ dS

t0 Q t t

dS dX Ω Q’ REMARK
dx
ds The sub-indexes (●)T and
(●)t are often dropped. But
Ω0 P P’
X x one must bear in mind that
stretch and unit elongation
always have a particular
direction associated to them.

35
Relation between Stretch and Unit
Elongation
 The stretch and unit elongation for a same point and direction are
related through:
ds − dS ds
ε= = − 1= λ − 1 ( −1 < ε < ∞ )
dS dS

 = (ε 0 ) =
If λ 1= ds dS : P and Q may have moved in time but have kept
the distance between them constant.

 If λ > 1 ( ε > 0 ) ds > dS : the distance between them P and Q has


increased with the deformation of the medium.

 If λ < 1 ( ε < 0 ) ds < dS : the distance between them P and Q has


decreased with the deformation of the medium.

36
Stretch and Unit Elongation in terms of
the Strain Tensors
 Considering:
( ds ) − ( dS )= 2 dX ⋅ E ⋅ dX
2 2
dX = T dS
( ds ) − ( dS )= 2 dx ⋅ e ⋅ dx dx = t ds
2 2

 Then:
× 1
( dS ) 2 = λ2
2
 ds 
( ds ) − ( dS ) = 2 ( dS ) T ⋅ E ⋅ T   − 1= 2 T ⋅ E ⋅ T
2 2 2

 dS  REMARK
2
 dS  E(X,t) and e(x,t)
( ds ) − ( dS ) = 2 ( ds ) t ⋅ e ⋅ t 1−   = 2 t ⋅e ⋅ t
2 2 2

 ds  contain information
× 1
( λ)
2
( ds )2 = 1 regarding the stretch
and unit elongation for
1
λ= any direction in the
λ = 1+ 2 T ⋅ E ⋅ T 1− 2 t ⋅e ⋅ t differential neighbour-
ε = λ −1 = 1 + 2 T ⋅ E ⋅ T −1 1 hood of a point.
ε = λ −1 = −1
1− 2 t ⋅e ⋅ t
37
2.6 Variation of Angles
Ch.2. Deformation and Strain

38
Variation of Angles
dx( ) = t ( ) ds ( )
1 1 1
(1) (1) (1)
dX = T dS
dx( ) = t ( ) ds (
2 2 2)
( 2) ( 2) ( 2) T(1)
dX =T dS t
t0
Q t(1)
t(2)
T(2)
Θ Ω R’ θ
dS(1) Q’
R ds(2) ds(1)
dS(2)
Ω0 P P’
X x

 The scalar product of the vectors dx(1) and dx(2) :

dx(1) ⋅ dx( 2) = dx(1) ⋅ dx( 2) cos θ = ds (1) ds ( 2) cos θ

39
Variation of Angles
d x( 2) =
x(1) ⋅ d ds (1) ds ( 2) cos θ
 dx(1) 
T
 dx ( 2 ) 
   

(1)
dx = F ⋅ dX (1) T(1) ⋅ (1 + 2E ) ⋅ T( 2)
cos θ =
 ( 2) ( 2) 1 + 2 T(1) ⋅ E ⋅ T(1) 1 + 2 T( 2 ) ⋅ E ⋅ T( 2 )
dx = F ⋅ dX
=  F ⋅ dX (1)  ⋅  F ⋅ dX (2)  = dX (1) ⋅ ( FT ⋅ F ) ⋅ dX (2)
T
(1) (2)
dx ⋅ dx
 
2E+1
dX(1) = T(1) dS (1) 

dX( ) = T( ) dS ( ) 
2 2 2

 1 1 
dx (1)=
⋅ dx (2)  T ⋅ ( 2E + 1 ) ⋅ T = ds (1)ds (2)  (1) (2) T(1) ⋅ ( 2E + 1) ⋅=
(1) (1) (2) ( 2)
dS 
dS T(2)  ds (1)ds (2) cos θ
ds(1) ds(2) λ λ 
(1) (2)
λ λ
λ = 1+ 2 T ⋅ E ⋅ T

 T(1) ⋅ (1 + 2E ) ⋅ T( 2) 
 
(1) (1)
 1+ 2 T ⋅ E ⋅ T 1 + 2 T( 2 ) ⋅ E ⋅ T( 2 ) 
40
Variation of Angles

dX(1) = T(1) dS (1) dx(1) = t (1) ds (1)


 ( 2)  ( 2) ( 2) ( 2)
( 2) ( 2) T(1) dx = t ds
dX = T dS t
t 0
Q t(1)
t(2)
T(2)
Θ Ω R’ θ
dS(1) Q’
R ds(2) ds(1)
dS(2)
Ω0 P P’
X x

 The scalar product of the vectors dX(1) and dX(2) :


X( 2) dX(1) ⋅ dX( 2) cos
dX(1) ⋅ d= = Θ dS (1) dS ( 2) cos Θ

41
Variation of Angles
X(1) =
d X( 2) dS (1) dS ( 2) cos Θ
⋅ d
 dX(1) 
T
 dX( 2) 
   
t (1) ⋅ (1 − 2e ) ⋅ t ( 2)
dX=(1)
F −1 ⋅ dx(1) cos Θ =
 ( 2) −1
dX= F ⋅ dx
( 2) 1 − 2 t (1) ⋅ e ⋅ t (1) 1 − 2 t ( 2) ⋅ e ⋅ t ( 2)
dX (1) ⋅ dX (2) =  F −1 ⋅ dx (1)  ⋅  F −1 ⋅ dx (2)  = dx (1) ⋅ ( F −T ⋅ F −1 ) ⋅ dx (2)
T


1−2e
dx(1) = t (1) ds (1)
 ( 2) ( 2) ( 2)
dx = t ds

dX (1) ⋅ dX (2)= (1) (1) (2)


 t ⋅ (1 - 2e ) ⋅ t ds
ds
(1) (1)
( 2) (1) (2) (1) ( 2 ) (1)
(2) (2)
(2)
((1) (2)
 = dS dS λ λ t ⋅ (1 - 2e ) ⋅ t = dS dS cos Θ )
λ dS λ dS

REMARK λ=
1
E(X,t) and e(x,t) contain information 1− 2 t ⋅e ⋅ t
regarding the variation in angles  t (1) ⋅ (1 − 2e ) ⋅ t ( 2) 
between segments in the differential  
(1) (1)
neighbourhood of a point.  1− 2 t ⋅e ⋅ t 1 − 2 t ( 2) ⋅ e ⋅ t ( 2) 
42
Example
Let us consider the motion of a continuum body such that the spatial description
of the Cartesian components of the spatial Almansi strain tensor is given by,
 0 0 −tetz 
 
[ e ( x, t ) ] =  0 0 0 
 tz 
 −te 0 t ( 2etz − et ) 

Compute at time t=0 (the reference time), the length of the curve that at time
t=2 is a straight line going from point a (0,0,0) to point b (1,1,1).
The length of the curve at time t=0 can be expressed as,
1
∫ λ ( x, t ) ds
B b
=L ∫=
A 
dS
a
= ds
λ
43
Example - Solution
The inverse of the stretch, at the points belonging to the straight line going
from a(0,0,0) to b(1,1,1) along the unit vector in the direction of the straight
line, is given by,
1
λ ( x, t ) = → λ −1 ( x, t ) = 1 − 2t ⋅ e ( x, t ) ⋅ t
1 − 2t ⋅ e ( x, t ) ⋅ t
Where the unit vector is given by,
1
[t ] = [1 1 1]
T

3
Substituting the unit vector and spatial Almansi strain tensor into the expression
of the inverse of the stretching yields,
2 t
λ −1
( x, t=) 1 + te
3

44
Example - Solution
The inverse of the stretch, which is uniform and therefore does not depends on
the spatial coordinates, at time t=2 reads,
4
λ −1 ( x, 2=
) 1 + e2
3

Substituting the inverse of the stretch into the integral expression provides the
length at time t=0,
4 2 4 2 (1,1,1)
( )
b b
∫Γ ∫a λ ∫a 3 ∫
−1
L= dS = x , 2 ds = 1 + e ds =1 + e ds =+
3 4e 2

3  
(0,0,0)

= 3

45
2.7 Physical Interpretation of E and e
Ch.2. Deformation and Strain

46
Physical Interpretation of E
 Consider the components of the material strain tensor, E:
 E XX E XY E XZ   E11 E12 E13 
= 
E = E XY EYY EYZ  E E22 E23 
 12
 E XZ EYZ EZZ   E13 E23 E33 

 For a segment parallel to the X-axis, the stretch is:


λ = 1+ 2 T ⋅ E ⋅ T
 E11 E12 E13  1 
reference T(1) ⋅ E
= ⋅ T(1) [1 0 0] ⋅  E12 E22 E23  ⋅=0 E
      11
configuration
 E13 E23 E33  0
T
T(1) 
  
T(1) 
 

Stretching of
1  dS  λ= 1 + 2E11 the material in
    1
T (1)
≡ 0  dX ≡ dS T (1)
=
0
0  0
the X-direction
   

47
Physical Interpretation of E
 Similarly, the stretching of the material in the Y-direction and the Z-
direction:
λ1 = 1 + 2 E11 ε X = λ X − 1= 1 + 2 E XX − 1
λ2 = 1 + 2 E22 ε Y = λY − 1= 1 + 2 EYY − 1
λ3 = 1 + 2 E33 ε Z = λZ − 1= 1 + 2 EZZ − 1

 The longitudinal strains contain information on the stretch and unit


elongation of the segments initially oriented in the X, Y and Z-directions
(in the material configuration).
 E XX E XY E XZ 
If E XX 0=
= εX 0 No elongation in the X-direction
E =  E XY EYY EYZ 
 E XZ EYZ EZZ  If EYY 0=
= εY 0 No elongation in the Y-direction
=
If EZZ 0=εZ 0 No elongation in the Z-direction

48
Physical Interpretation of E
 Consider the angle between a segment parallel to the X-axis and a
segment parallel to the Y-axis, the angle is:
T ( ) ⋅ ( 1 + 2E ) ⋅ T ( )
1 2

cos θ =
1 + 2 T(1) ⋅ E ⋅ T(1) 1 + 2 T( 2 ) ⋅ E ⋅ T( 2 )

reference 1  0  T(1) ⋅ T( 2) =
0
    T ⋅ E ⋅ T(1) =
(1)
configuration T(1) = 0  T( 2) = 1  E11
0 
 
0 
  T(1) ⋅ E ⋅ T( 2) =E12
( 2) ( 2)
T ⋅E ⋅T = E22
2 E12
cos θ =
1 + 2 E11 1 + 2 E 22

deformed
configuration

2 E XY π 2 E XY
θ ≡ θ xy =arccos = − arcsin
1 + 2 EXX 1 + 2 EYY 2 1 + 2 EXX 1 + 2 EYY

49
Physical Interpretation of E
π 2 E XY
θ ≡ θ xy = − arcsin
2 1 + 2 EXX 1 + 2 EYY

 The increment of the final angle w.r.t. its initial value:

2 E XY
θ xy − Θ
∆Θ XY =

=− arcsin
1 + 2 EXX 1 + 2 EYY
XY
π
2
reference
configuration

deformed
configuration

50
Physical Interpretation of E
 Similarly, the increment of the final angle w.r.t. its initial value for
couples of segments oriented in the direction of the coordinate axes:
2 E XY
∆Θ XY = − arcsin
1 + 2 EXX 1 + 2 EYY
2 E XZ
∆Θ XZ = − arcsin
1 + 2 EXX 1 + 2 EZZ
2 EYZ
∆ΘYZ = − arcsin
1 + 2 EYY 1 + 2 EZZ

 The angular strains contain information on the variation of the angles


between segments initially oriented in the X, Y and Z-directions (in the
material configuration). If E = 0 No angle variation between the
XY
 E XX E XY E XZ  X- and Y-directions
E =  E XY EYY EYZ  If E XZ = 0 No angle variation between the
 E XZ EYZ EZZ  X- and Z-directions
If EYZ = 0 No angle variation between the
Y- and Z-directions
51
Physical Interpretation of E
 In short,
deformed
configuration

reference λ1 dX= 1 + 2 E XX dX
configuration
λ2 dY= 1 + 2 EYY dY
λ3 dZ= 1 + 2 EZZ dZ

2 E XY
∆Θ XY = − arcsin
1 + 2 EXX 1 + 2 EYY
2 E XZ
∆Θ XZ = − arcsin
1 + 2 EXX 1 + 2 EZZ
2 EYZ
∆ΘYZ = − arcsin
1 + 2 EYY 1 + 2 EZZ

52
Physical Interpretation of e
 Consider the components of the spatial strain tensor, e:
 exx exy exz   e11 e12 e13 
   
e ≡ exy eyy eyz  =  e12 e22 e23 
 exz eyz ezz   e13 e23 e33 

 For a segment parallel to the x-axis, the stretch is:


1
λ=
1− 2 t ⋅e ⋅ t
deformed  e11 e12 e13  1 
configuration t ⋅ e ⋅ t [1 0 0] ⋅ e12
= e22 e23 =
⋅ 0  e11
 e13 e23 e33  0 

1  ds  1 Stretching of
    λ1 = the material in
t (1) ≡ 0  dx ≡  0  1 − 2e11
0  0 the x-direction
   
53
Physical Interpretation of e
 Similarly, the stretching of the material in the y-direction and the z-
direction: 1 1
λ1 = ⇒ ε x = λx − 1 = −1
1 − 2e11 1 − 2exx
1 1
λ2 = ⇒ ε y = λy − 1= −1
1 − 2e22 1 − 2eyy
1 1
λ3 = ⇒ ε z = λz − 1 = −1
1 − 2e33 1 − 2ezz

 The longitudinal strains contain information on the stretch and unit


elongation of the segments oriented in the x, y and z-directions (in the
deformed or actual configuration).
 exx exy exz 
 
e ≡ exy eyy eyz 
 exz eyz ezz 

54
Physical Interpretation of e
 Consider the angle between a segment parallel to the x-axis and a
segment parallel to the y-axis, the angle is:
t (1) ⋅ (1 − 2e ) ⋅ t ( 2)
cos Θ =
reference deformed 1 − 2 t (1) ⋅ e ⋅ t (1) 1 − 2 t ( 2) ⋅ e ⋅ t ( 2)
configuration configuration
1  0  t (1) ⋅ t ( 2) = 0
(1) (1)
    t ⋅ e ⋅t = e11
t (1) = 0  t (2) = 1 
0  0  t (1) ⋅ e ⋅ t ( 2) =e12
    ( 2)
t ⋅e ⋅t = ( 2)
e22
−2 e12
cos Θ =
1 − 2 e11 1 − 2 e 22

π 2exy
Θ ≡ Θ XY= + arcsin
2 1 − 2 exx 1 − 2 eyy

55
Physical Interpretation of e
π 2exy
Θ ≡ Θ XY= + arcsin
2 1 − 2 exx 1 − 2 eyy

 The increment of the angle in the reference configuration w.r.t. its value
in the deformed one:
2exy
∆θ xy = θ xy − Θ XY = − arcsin
 1 − 2 exx 1 − 2 eyy
π
2

reference deformed
configuration configuration

56
Physical Interpretation of e
 Similarly, the increment of the angle in the reference configuration w.r.t.
its value in the deformed one for couples of segments oriented in the
direction of the coordinate axes:
π 2exy
∆θ xy = − Θ XY = − arcsin
2 1 − 2 exx 1 − 2 eyy
π 2exz
∆θ xz = − Θ XZ = − arcsin
2 1 − 2 exx 1 − 2 ezz
π 2eyz
∆θ yz = − ΘYZ = − arcsin
2 1 − 2 eyy 1 − 2 ezz

 The angular strains contain information on the variation of the angles


between segments oriented in the x, y and z-directions (in the deformed
or actual configuration).  exx exy exz 
 
e ≡ exy eyy eyz 
 exz eyz ezz 

57
Physical Interpretation of e
 In short, reference
configuration
deformed
configuration
dx
= 1 − 2exx dx
λ1
dy
= 1 − 2eyy dy
λ2
dz
= 1 − 2ezz dz
λ3

π 2exy
∆θ xy = − Θ XY = − arcsin
2 1 − 2 exx 1 − 2 eyy
π 2exz
∆θ xz = − Θ XZ = − arcsin
2 1 − 2 exx 1 − 2 ezz
π 2eyz
∆θ yz = − ΘYZ = − arcsin
2 1 − 2 eyy 1 − 2 ezz

58
2.8 Polar Decomposition
Ch.2. Deformation and Strain

59
Polar Decomposition
 Polar Decomposition Theorem:
 “For any non-singular 2nd order tensor F there exist two unique
positive-definite symmetrical 2nd order tensors U and V, and a unique
orthogonal 2nd order tensor Q such that: ”
not
 left polar
=
U F ⋅F
T
 decomposition
not 
V = F ⋅ FT  F = Q ⋅ U = V ⋅Q
Q= F ⋅ U −1 =V −1 ⋅ F  right polar
 decomposition
 The decomposition is unique. REMARK
 Q: Rotation tensor An orthogonal 2nd
order tensor verifies:
 U: Right or material stretch tensor
QT ⋅ Q = Q ⋅ QT = 1
 V: Left or spatial stretch tensor

60
Properties of an orthogonal tensor
 An orthogonal tensor Q when multiplied (dot product) times
a vector rotates it (without changing its length): y= Q ⋅ x
 y has the same norm as x:
y = y ⋅ y = [ y ] [ y ] = [Q ⋅ x ] ⋅ [Q ⋅ x ] = x ⋅ Q
⋅Q⋅x = x
2 T T T 2

 when Q is applied on two vectors x(1) and x(2), with the same origin,
the original angle they form is maintained:
T T
 (1)   (2) 
 y   y 

  
y = Q⋅x
(1) (1)
y ⋅y
(1)
x ⋅ QT ⋅ Q ⋅ x (2)
(2)
(1)
x (1) ⋅ x (2)
= = = cos α
y (2)= Q ⋅ x (2) y (1) y (2) y (1) y (2) x (1) x (2)

 Consequently, the rotation y= Q ⋅ x maintains angles and


distances.

61
Polar Decomposition of F
 Consider the deformation gradient tensor, F:

F = Q ⋅ U = V ⋅Q
stretching
 
rotation
 
dx =F ⋅ dX =( V ⋅ Q ) ⋅ dX =V ⋅ ( Q ⋅ dX )

(not)
F (•) ≡ stretching  rotation (•)

rotation

stretching
 
dx =F ⋅ dX =( Q ⋅ U ) ⋅ dX =Q ⋅ ( U ⋅ dX )
REMARK
For a rigid body motion:
not
F(•) ≡ rotation  stretching (•) U= V= 1 and Q = F

62
2.9 Volume Variation
Ch.2. Deformation and Strain

63
Differential Volume Ratio
 Consider the variation of a differential volume associated to a particle
P:
=
dV0 ( dX
(1) ( 2)
× dX (3)
)=
⋅ dX
 dX 1(1) dX 2(1) dX 3(1) 
deformed  ( 2) ( 2) ( 2)

configuration = det
=  dX 1 dX 2 dX 3  M
 ( 3) ( 3) ( 3) 

dX 1 dX 2

dX 3 

M 
reference  

configuration
=
dVt ( dx( ) × dx( ) )=
1 2
⋅ dx( )
3

 dx1(1) dx2(1) dx3(1) 


 ( 2) 
= det
=  dx1 dx2( 2) dx3( 2)  m
 ( 3) 
 dx1 dx2( 3) dx3( 3) 

 m 

= =
M ij dX (i )
j mij dx (ji )

64
Differential Volume Ratio
 Consider now: dx (i ) =
F ⋅ dX ( i ) i ∈ {1, 2,3} → Fundamental eq. of deformation
 (i )
dx j = Fjk ⋅ dX k(i ) i, j ∈ {1, 2,3}

= =
M ij dX (i )
j and mij dx (ji )

m=
ij dx (j=
i)
Fjk dX k(=
i)
Fjk M ik= M ik FkjT = M ⋅ FT
m

 Then:
dVt = m = M ⋅ FT = M FT = F M = F dV0 dVt = F dV0

dV
0

 And, defining J (X,t) as the jacobian of the deformation,


J ( X, t ) det F ( X, t ) > 0
= dVt= J ⋅ dV0

65
2.10 Area Variation
Ch.2. Deformation and Strain

66
Surface Area Ratio
 Consider the variation of a differential area associated to a particle P:
dA := dAN → material vector "differential of area" → dA = dA
da := dan → spatial vector "differential of area" → da = dA
Deformed (current)
configuration ( 3)
dV0 = dH dA = d
X
 ⋅
N dA =
Reference (initial) dH
configuration =dX(3) ⋅ N
 dA =d A ⋅ d X ( 3)

dA

dVt = dh da = x( ) ⋅ n da =
3
d
dh
=dx ( 3 ) ⋅ n
 da =d a ⋅ d x ( 3)

da

67
Surface Area Ratio

 Consider now:
da ⋅ dx (
3)
dV=
t

dx (3)= F ⋅ dX( )
3

dVt = F dV0
dV= dA ⋅ dX( )
3
0

( 3) ( 3) ( 3)
dVt = F d
A ⋅
dX =
  d a ⋅ F ⋅ dX ∀dX ⇒ F dA = da ⋅ F

F dV dV0 dVt
0

da = F ⋅ dA ⋅ F −1 =
da n F N ⋅ F −1dA da F N ⋅ F −1 dA
=
dA = N dA
da = n da
68
2.11 Volumetric Strain
Ch.2. Deformation and Strain

69
Volumetric Strain
 Volumetric Strain:
def dV ( X, t ) − dV ( X, t0 ) not dVt − dV0
e ( X, t ) =
dV ( X, t ) dV0

dVt = F dV0

F dV0 − dV0
e=
dV0

=
e F −1 REMARK
The incompressibility condition (null
volumetric strain) takes the form
e = J −1 = 0 ⇒ J = F = 1

70
2.12 Infinitesimal Strain
Ch.2. Deformation and Strain

71
Infinitesimal Strain Theory
 The infinitesimal strain theory (also called t0 t
small strain theory) is based on the Ω
simplifying hypotheses: P u P’
Ω0 X
x
 Displacements are very small w.r.t. the
typical dimensions in the continuum
medium,

 As a consequence, u << ( size of Ω0 ) and the reference and deformed configurations are
considered to be practically the same, as are the material
not
and spatial coordinates:
Ω ≅ Ω0 and x = X+u ≅ X ( X, t ) u ( X, t ) ≡ u ( x, t )
U=
xi = X i + ui ≅ X i not
U i ( X, t ) =ui ( X, t ) ≡ ui ( x, t ) i ∈ {1, 2,3}

Displacement gradients are infinitesimal, ∂ui


 << 1, ∀i, j ∈ {1, 2,3}.
∂x j

72
Infinitesimal Strain Theory
 The material and spatial coordinates coincide, x = X + 
u ≅X
≈0
 Even though it is considered that u cannot be neglected when calculating
other properties such as the infinitesimal strain tensor ε.
 There is no difference between the material and spatial
differential operators:

symb
∂ ∂
∇ = ˆ
e = eˆ i =

∂ ∂
i
 X i x i
J (=
 X, t ) U( X, t )= ⊗ ∇ u(x, t )=
⊗ ∇ j(x, t )

 The local an material time derivatives coincide


Γ ( X (x, t ) γ =
 , t ) ≈ Γ= (x, t ) γ ( X, t )
 ≈x
 d γ ∂γ ( X, t ) ∂γ ( x, t )

= = = γ
 dt ∂t ∂t
73
Strain Tensors
 Green-Lagrange strain tensor

E= (
1 T
F F − 1)=
1
( J + JT + JT J )  E ≅
1
2
( J + J T
) =
1
2
( j + jT
)= ε
2 2 

1  ∂ui ∂u j ∂uk ∂uk   Eij ≅ = 1  ∂ui ∂u j 
E=  + +  ε ij  + =  ε ij i, j ∈ {1, 2,3}
ij
2  ∂x j ∂xi ∂xi ∂xi   2  ∂x j ∂xi 
∂uk
<< 1
∂x j
 Euler-Almansi strain tensor 
 e ≅
1
2
( j + jT
) =
1
2
( J + JT ) = ε

e=
1
( 1 − F −T F −1 ) =
1
( j + jT − jT j)   ∂u 
2 2  eij ≅ 1  ∂ui + j =  ε ij i, j ∈ {1, 2,3}
1  ∂ui ∂u j ∂uk ∂uk   2  ∂x j ∂xi 
e=  + − 
2  ∂x j ∂xi ∂xi ∂xi  ∂uk
ij
<< 1
∂x j
 Therefore, the infinitesimal strain tensor is defined as :
 REMARK
1
( ) 1
( ) 1 not

=ε J + J=T
j + j=
T
(u ⊗ ∇ + ∇ ⊗ =
u ) ∇ su
 2 2 2 ε is a symmetrical tensor and its components

 ε ij =1  ∂ui ∂u j  are infinitesimal: | ε ij |<< 1, ∀i, j ∈ {1, 2,3}
 +  i, j ∈ {1, 2,3}
 2  ∂x j ∂xi 
74
Stretch and Unit Elongation
 Stretch in terms of the strain tensors:
1
λT = 1 + 2 
T⋅⋅T
E  λt =
x 1− 2 t ⋅e ⋅ t

Considering that e ≅ E ≅ ε and that it is


infinitesimal, a Taylor linear series expansion up to
first order terms around x = 0 yields:
λ ( x=
) 1+ 2 x λ ( x) =
1
dλ 1− 2 x
≅ λ ( 0) + x=
1+ x

dx x =0
 ≅ λ ( 0) + x=
1+ x
=1 dx x =0

λT ≅ 1 + T ⋅ ε ⋅ T =1 λt ≅ 1 + t ⋅ ε ⋅ t

 But in Infinitesimal Strain Theory, T ≈ t. So the linearized stretch and


unit elongation through a direction given by the unit vector T ≈ t are:
ds ds − dS
λ= ≅ 1+ t ⋅ε ⋅t ≅ 1+ T ⋅ε ⋅ T ε= = λ −1 = t ⋅ ε ⋅ t
dS dS
75
Physical Interpretation of
Infinitesimal Strains
 Consider the components of the infinitesimal strain tensor, ε:
ε xx ε xy ε xz  ε11 ε12 ε13 
 
=ε ε xy ε yy ε yz  ≡ ε12 ε 22 ε 23 
ε xz ε yz ε zz  ε13 ε 23 ε 33 

 For a segment parallel to the x-axis, the stretch and unit elongation are:
λ ≅ 1+ t ⋅ε ⋅t
ε11 ε12 ε13  1 
reference t ⋅ ε ⋅ t [1 0 0] ⋅ ε12 ε 22 ε 23  =
= ⋅ 0  ε11
configuration ε13 ε 23 ε 33  0 

λ1 = λx = 1 + ε11 Stretch in the x-direction


1  ds  ε1 =1 − λ =ε11
    Unit elongation in the
T (1)
≅t (1)
≡  0  dX ≅ dx ≡  0  ε x =1 − λ =ε xx x-direction
0  0
   
76
Physical Interpretation of
Infinitesimal Strains
 Similarly, the stretching and unit elongation of the material in the y-
direction and the z-direction:
λ1 = 1 + ε11 ⇒ ε x = λx − 1 = ε xx
λ2 = 1 + ε 22 ⇒ ε y = λ y − 1 = ε yy
λ3 = 1 + ε 33 ⇒ ε z = λz − 1 = ε zz

 The diagonal components of the infinitesimal strain tensor are the unit
elongations of the material when in the x, y and z-directions.
ε xx ε xy ε xz 
 
ε = ε xy ε yy ε yz 
ε xz ε yz ε zz 

77
Physical Interpretation of
Infinitesimal Strains
 Consider the angle between a segment parallel to the X-axis and a
segment parallel to the Y-axis, the angle is Θ = π2 . XY

 Applying: π 2E
θ ≡ θ xy = − arcsin XY
2 1 + 2 EXX 1 + 2 EYY
E XX = ε xx
E XY = ε xy
EYY = ε yy

π 2ε xy π π
reference
θ xy = − arcsin ≅ − arcsin 2ε xy = − 2ε xy
2 1 + 2ε xx 1 + 2ε yy 2  2
configuration     ≈2ε xy
≈1 ≈1

REMARK
The Taylor linear series expansion of arcsin x yields
( x ) + ... = x + O ( x 2 )
d arcsin
arcsin ( x ) ≅ arcsin ( 0 ) +
dx x =0

78
Physical Interpretation of
Infinitesimal Strains
π
θ xy ≅ − 2ε xy
2

 The increment of the final angle w.r.t. its initial value:


π π π
∆θ xy =
θ xy − ≅ − 2ε xy − =−2ε xy
2 2 2
 Similarly, the increment of the final angle w.r.t. its initial value for
couples of segments oriented in the direction of the coordinate axes:
1 1 1
ε xy =− ∆θ xy ; ε xz =− ∆θ xz ; ε yz =− ∆θ yz
2 2 2
 The non-diagonal components of the infinitesimal strain tensor are equal
to the semi-decrements produced by the deformation of the angles
between segments initially oriented in the x, y and z-directions.

79
Physical Interpretation of
Infinitesimal Strains
 In short,
reference deformed
configuration configuration

ε xx = ε x 1
ε xy =− ∆θ xy
2
ε yy = ε y 1
ε xz =− ∆θ xz
2
ε zz = ε z 1
ε yz =− ∆θ yz
2

80
Engineering Strains
 Using an engineering notation, instead of the scientific notation, the
components of the infinitesimal strain tensor are
REMARK
Positive longitudinal strains indicate
increase in segment length.

Positive angular strains indicate the


corresponding angles decrease with
the deformation process.
Angular
strains Longitudinal
strains

 Because of the symmetry of ε, the tensor can be written as a


6-component infinitesimal strain vector, (Voigt’s notation):
def
ε∈R 6
[ε x , ε y , ε z ,
ε= γ xy , γ xz , γ yz ]T
   
longitudinal angular strains
strains
81
Variation of Angles
 Consider two segments in the reference configuration with the same
origin an angle Θ between them.
θ = Θ + ∆θ
T ⋅ (1 + 2E ) ⋅ T
(1) ( 2)
E=ε
cos θ =
(1) (1) ( 2) ( 2)
1+ 2 T ⋅E ⋅T 1+ 2 T ⋅E ⋅T
T(1) ⋅ [1 + 2ε ] ⋅ T(2)
cos(Θ + ∆θ ) =
1 + 2T(1) ⋅ ε ⋅ T(1) 1 + 2T(2) ⋅ ε ⋅ T(2)
reference deformed
configuration ≈1 ≈1
configuration

cos(Θ + ∆θ ) =
T(1) ⋅ T(2) + 2T(1) ⋅ ε ⋅ T(2)

82
Variation of Angles
cos(Θ + ∆θ=) T(1) ⋅ T(2) + 2T(1) ⋅ ε ⋅ T(2)

 T(1) and T(2) are unit vectors in the directions of the original segments,
therefore, T(1) ⋅ T(2)= T(1) T(2) cos Θ
= cos Θ
T(1) ≈ t (1)
 Also, cos(Θ + ∆θ=) cos Θ ⋅ cos ∆θ − sinΘ ⋅ sin∆=
θ cos Θ − sinΘ ⋅ ∆θ T (2)
≈ t (2)

Θ ≈θ
≈1 ≈ ∆θ
2T(1) ⋅ ε ⋅ T(2) 2t (1) ⋅ ε ⋅ t (2)
θ cos Θ + 2T ⋅ ε ⋅ T
cos Θ − sinΘ ⋅ ∆= (1) (2)
∆θ =− =−
sinΘ sinθ

REMARK
The Taylor linear series expansion of sin x
and cos x yield
( x ) + ... = x + O ( x 2 )
d sin
sin ( x ) ≅ sin ( 0 ) +
dx x =0

( x ) + ... =1 + O ( x 2 )
d cos
cos ( x ) ≅ cos ( 0 ) +
dx x =0

83
Polar Decomposition
 Polar decomposition in finite-strain problems:
left polar
not
 decomposition
=
U FT ⋅ F 
not 
V = F ⋅ FT  ⇒ F = Q ⋅ U = V ⋅Q
Q= F ⋅ U −1 =V −1 ⋅ F  right polar
 decomposition

REMARK
In Infinitesimal Strain Theory
x ≈ X , therefore,= ∂x
F ≈1
∂X

84
Polar Decomposition
 In Infinitesimal Strain Theory:
U = FT F = (1 + J ) ⋅ ( 1 + J ) =
T
1 + J + JT + JT ⋅ J ≈ 1 + J + JT = 1 +
1
2
( J + JT )
<< J =x

U= 1 + ε
infinitesimal strain tensor

Similarly, REMARK
The Taylor linear series expansion of 1 + x
U = (1 + ε ) = 1 − ε = 1 − ( J + J T )
−1 1
and (1 + x ) yield
−1 −1

=x 2

x =1 + x + O ( x 2 )
1
λ ( x ) = 1 + x ≅ λ ( 0) +
=ε dx x =0 2
U −1 = 1 − ε λ ( x ) = (1 + x ) ≅ λ ( 0 ) +
−1 dλ
x =1 − x + O ( x 2 )
dx
infinitesimal x =0

strain tensor
85
Polar Decomposition
(1 + J ) ⋅ 1 − ( J + JT ) =+
1 J − ( J + J T ) − J ⋅ ( J + J T ) =+ ( J − JT )
1 1 1 1
F U −1 =
Q =⋅ 1
 2  2 2 2
<< J = Ω
Q= 1 + Ω
 The infinitesimal rotation tensor Ω is defined: REMARK
The antisymmetric or
 def 1 1 def
Ω = (J − J T )= (u ⊗ ∇ − ∇ ⊗ u)= ∇ a u skew-symmetrical
 2 2 gradient operator is

Ω 1  ∂ui ∂u j  defined as:
=  −  << 1 i, j ∈ {1, 2,3}
 ij 2  ∂x j ∂xi  1
  0 Ω12 −Ω31 
∇ a (•=
) [(•) ⊗ ∇ − ∇ ⊗ (•)]
2
 The diagonal terms of Ω are zero: [ Ω ] =  −Ω12 0 Ω 23 
 Ω31 −Ω 23 0 
 It can be expressed as
an infinitesimal rotation vector θ,  ∂u3 ∂u2 
 ∂x − ∂x 
θ1   −Ω 23   2 3 
REMARK
    1  ∂u1 ∂u3  def 1
θ ≡ θ 2 =  −Ω31 =  − = ∇×u Ω is a skew-symmetric
θ   −Ω  2  ∂x3 ∂x1  2 tensor and its components
 3   12   ∂u2 ∂u1 
 −  are infinitesimal.
 ∂x1 ∂x2 

86
Polar Decomposition
 From any skew-symmetric tensor Ω, it can be extracted a vector θ (axial
vector of Ω) exhibiting the following property:
Ω ⋅r =θ×r ∀r
As a consequence:
 The resulting vector is orthogonal to r.

 If the components of Ω are infinitesimal, then Ω ⋅ r =θ × r is also infinitesimal


 The vector r + Ω ⋅ r = r + θ × r can be seen as the result of applying a (infinitesimal)
rotation (of axial vector θ) on the vector r .

87
Proof of θ×r =
Ω⋅r ∀r

 The result of the dot product of the infinitesimal rotation tensor, Ω, and a
generic vector, r, is exactly the same as the result of the cross product of
the infinitesimal rotation vector, θ, and this same vector.
 0 Ω12 −Ω31  θ1   −Ω 23   r1 
[Ω ] =  −Ω12 0 Ω23  → θ ≡ θ 2  =  −Ω31  ⇒ Ω ⋅ r =θ × r ∀r =  r2 
θ   −Ω  r 
 Ω31 −Ω 23 0   3   12   3
 Proof:
eˆ1 eˆ 2 eˆ 3   eˆ1 eˆ 2 eˆ 3   Ω12 r2 − Ω31r3 
 
θ × r = det θ1 θ 2 θ3  = det  −Ω 23 −Ω12  =
not
−Ω31 −Ω12 r1 + Ω 23 r3 
 r1 r2 r3   Ω r −Ω r 
 r1 r2 r3   31 1 23 2 

 0 Ω12 −Ω31   r1   Ω12 r2 − Ω31r3 


    
Ω ⋅ r =−Ω12
 0 Ω 23  r2 = −Ω12 r1 + Ω 23 r3 
 Ω31 −Ω 23 0   r3  
 Ω31r1 − Ω 23 r2 

88
Polar Decomposition
 Using:
J= F − 1


1
2
( J + JT ) 1 J =+
F =+ 1
1
2
( J + JT ) + ( J − JT )
1
2 F = 1+ε+Ω
Q= 1 + Ω =ε = Ω

 Consider a differential segment dX:


stretch
 rotation

dx = F ⋅ dX = ( 1 + ε + Ω ) ⋅ d X = ε ⋅ d X + ( 1 + Ω ) ⋅ dX

F(•) ≡ stretching (•) + rotation (•)

REMARK
The infinitesimal rotation tensor
characterizes the rotation and, in the
small-strain context, maintains angles
89 and distances.
Volumetric Deformation
 The volumetric strain:
=
e F −1

 Considering: F= Q ⋅ U and U= 1 + ε

1 + ε xx ε xy ε xz
F = Q ⋅ U = Q U = U = 1 + ε = det ε xy 1 + ε yy ε yz =
ε xz ε yz 1 + ε zz
= 1 + ε xx + ε yy + ε zz + O ( ε 2 ) ≈ 1 + Tr ( ε )
= Tr ( ε )

e = Tr ( ε )

90
2.13 Strain Rate
Ch.2. Deformation and Strain

REMARK
We are no longer assuming an
infinitesimal strain framework
91
Spatial Velocity Gradient Tensor
 Consider the relative velocity between two points in space at a given
(current) instant:
∂v
=
 v P′ v= (x, t ) v ( x1 , x2 , x3 , t ) dv = ⋅ dx =⋅
l dx
 ∂
x
dv (x, t ) = v Q′ − v P′ = v ( x + dx, t ) − v ( x, t ) l
∂vi
=
dvi = dx j lij dx j
∂x j

lij i, j ∈ {1, 2,3}

 ( )= v ⊗ ∇
def ∂v x, t

 l ( x , t )=
Spatial velocity  ∂x

gradient tensor = lij ∂vi i, j ∈ {1, 2,3}
 ∂x j

92
Strain Rate and Rotation Rate (or Spin)
Tensors
 The spatial velocity gradient tensor can be split into a symmetrical and
a skew-symmetrical tensor:

l= v ⊗ ∇
 sym [ l ] + skew [ l ] =
l= :d + w
l ∂vi i, j ∈ 1, 2,3
=
 ij ∂x { }
 j

Strain Rate Tensor Rotation Rate or Spin Tensor


( l + l )=
def
1 1 not
d= sym(l)= T
( v ⊗ ∇ + ∇ ⊗ v )= ∇ s v def
w= skew (l)=
1
( l − l )=
T 1 not
( v ⊗ ∇ − ∇ ⊗ v )= ∇ a v
2 2 2 2
1  ∂v ∂v  1  ∂v ∂v 
d ij = i + j  i, j ∈ {1, 2,3} w ij = i − j  i, j ∈ {1, 2,3}
2  ∂x j ∂xi  2  ∂x j ∂xi 
 d11 d12 d 31   0 w12 − w 31 
[d ] = d12 d 22 d 23  [ w ] =  − w12 0 w 23 
d 31 d 23 d 33   w 31 − w 23 0 

93
Physical Interpretation of d
 The strain rate measures the rate of deformation of the square of the
differential length ds in the spatial configuration,
d d d d  dx   dx 
( ds(t =
)) ( dx ⋅ d=x) ( dx ) ⋅ dx + dx ⋅ ( d=
x ) d   ⋅ dx + dx ⋅ d  =  dv ⋅ dx + dx ⋅ dv
2

dt dt dt dt  
dt  
dt
dv = l ⋅ dx =v =v
1
= d (l + lT )
2
d
dt
( ds(t ) )
2
= (
d x ⋅ l ) ⋅ d x + d x ⋅ ( l ⋅ dx ) =
 
T

dx ⋅  lT + l  ⋅ dx = 2dx ⋅ d ⋅ dx
dv 
T dv  = 2d
   

Differentiating w.r.t. time the expression ( ds(t ) ) − ( dS )= 2 dX ⋅ E ⋅ dX


=
2 2

d
dt
( (ds (t )) 2 − (
dS ) 2 ) =
d
dt
( 2dX ⋅ E ( X, t ) ⋅ dX ) = 2dX ⋅
dE
dt
⋅ dX =
d
dt
( (ds (t )) 2 )
 constant
2dX⋅E X,t ⋅dX notation
  
= E

94
Physical Interpretation of d
 ⋅ dX = dx ⋅ d ⋅ dx
dX ⋅ E
dx= F ⋅ dX
 ⋅ dX = dx ⋅ d ⋅ dx =
dX ⋅ E [
dx ] [ d ][ dx ] = [ F ⋅ dX ] [d ][ F ⋅ dX ] =
T
   
T

 
dX ⋅ ( FT ⋅ d ⋅ F) ⋅ dX
F⋅dX F⋅dX T T
dX  F  F dX 
  

 And, rearranging terms:


  ⋅ dX= 0 ∀dX
dX ⋅ FT ⋅ d ⋅ F − E =
FT ⋅ d ⋅ F − E  = FT ⋅ d ⋅ F
E
  0

 There is a direct relation between the material derivative of the material strain
tensor and the strain rate tensor but they are not the same.
  and d will coincide when in the
E
reference configuration F |t =t0 = 1 .
REMARK
Given a 2 order tensor A,
nd

if x ⋅ A ⋅ x =0 for any vector


x ≠ 0 then A = 0 .

95
Physical Interpretation of w
 To determine the (skew-symmetric) rotation rate (spin) tensor only three
different components are needed:
 0 w12 w13 
1  ∂vi ∂v j 
w ij = −  i, j ∈ {1, 2,3} [ w ] =  − w12 0 w23 
2  ∂x j ∂xi   − w13 − w23 0 

 The spin vector (axial vector [w]) of can be extracted:


  ∂v 2 ∂v3  
−  − 
  3 ∂x ∂x2 
   − w23   ω1   0 −ω3 ω2 
1   ∂v3 ∂v1   w=  ω = ω −ω1 
1 1
ω
= rot ( =
v) ∇×v ≡ − − =
    2 [w]  3 0
2   ∂x1 ∂x3  
13
2 2
   − w12  ω3   −ω2 ω1 0 
  ∂v1 ∂v 2  
 −  ∂x − ∂x  
  2 1 

 The vector 2ω = ∇ × v is named vorticity vector.

96
Physical Interpretation of w
 It can be proven that the equality ω× r =w ⋅ r ∀r holds true.
Therefore:
 ω is the angular velocity of a rotation movement.
 ω x r = w · r is the rotation velocity of the point that
has r as its position vector w.r.t. the rotation centre.

 Consider now the relative velocity dv,


dv = l ⋅ dx
l= d + w

dv =d ⋅ dx + w ⋅ dx

97
2.14 Material time Derivatives
Ch.2. Deformation and Strain

99
Deformation Gradient Tensor F
 The material time derivative of the deformation gradient tensor,

∂xi ( X, t ) REMARK
=
Fij ( X, t ) i, j ∈ {1, 2,3} The equality of cross derivatives
∂X j
applies here: ∂ 2 (•) = ∂ 2 (•)
d ∂µi µ j ∂µ j µi
dt
dFij ∂ ∂xi ( X, t ) ∂ ∂xi ( X, t ) ∂Vi ( X, t ) ∂vi ( x ( X, t ) ) ∂x k
= = = = = lik Fkj
dt ∂t ∂X j ∂X j ∂t ∂X j ∂xk ∂X j
=Vi ( X, t ) =lik =Fkj

 dF notation 
 dt = F= l ⋅ F

 dF=ij
F= lik Fkj i, j ∈ {1, 2,3}
 dt ij

100
Inverse Deformation Gradient Tensor F-1
 The material time derivative of the inverse deformation gradient
tensor,
1
F ⋅ F −1 = REMARK
Do not mistake the material derivative
d of the inverse tensor for the inverse of
dt the material derivative of the tensor:
d ( F −1 )
( F ( X,t ) ) ≠ ( F
 ( X,t ) )
d −1 dF −1 d −1
(F ⋅ F ) = ⋅F + F⋅ = 0 −1
dt dt dt dt
d ( F −1 ) dF −1
⇒ F⋅ =− ⋅ F =−F ⋅ F −1
dt dt
Rearranging terms,
d ( F −1 )
−F −1 ⋅ F ⋅ F −1 =
= −F −1 ⋅ l ⋅ F ⋅ F −1 =
−F −1 ⋅ l
dt  d ( F −1 )
= l⋅F =1  = −F −1 ⋅ l
 dt
 −1
 dFij
 dt = − F −1
ik lkj i, j ∈ {1, 2,3}

101
Strain Tensor E
 The material time derivative of the material strain tensor has already
been derived for the physical interpretation of the deformation rate
tensor: 
E= F ⋅d ⋅F
T

 A more direct procedure yields the same result:

=
E
2
( F ⋅ F − 1)
1 T

F = l ⋅ F
d F=T
F T ⋅ lT
dt
dE  1  T
dt
= E=
2
( F ⋅ F + FT ⋅ F ) =
2
( F ⋅ l ⋅ F + FT ⋅ l ⋅ F ) = FT ⋅ ( l + lT ) ⋅ F = FT ⋅ d ⋅ F
1 T T 1
2
 
d

 = FT ⋅ d ⋅ F
E

102
Strain Tensor e
 The material time derivative of the spatial strain tensor,

e=
1
2
( 1 − F −T ⋅ F −1 )
F=−1
F −1 ⋅ l
d F −T= lT ⋅ F −T
dt

− ( F −T ⋅ F −1 + F −T ⋅ F −1 ) =( lT ⋅ F −T ⋅ F −1 + F −T ⋅ F −1 ⋅ l )
de 1 1
e =
=
dt 2 2

e = ( l ⋅ F ⋅ F + F −T ⋅ F −1 ⋅ l )
1 T −T −1
2

103
Volume differential dV
 The material time derivative of the volume differential associated to a
given particle,

dV ( x( X, t ), t ) = F( X, t ) dV0 ( X) The material time derivative of the determinant of the


deformation gradient tensor is:
d
dt For a 2nd order
d ∂ F( X, t ) d tensor A:  d A = d A= A ⋅ A −1
dV ( t ) = dV0 F dV0  
∂t
ji
dt dt  dA ij dAij

d F d F dFij −1 dFij −1
= = F F= F F F lik
d dt dFij dt
ji
dt 
kj ji
( dV=) ( ∇ ⋅ v ) F dV0 lik Fkj ( F⋅F−1 ) δ
= ki
dt
= dV ki

∂v i dF
= F=
lii F = F ∇⋅v = F ∇ ⋅ v = (∇ ⋅ v ) F
d ∂xi
 dt
( dV (x, t ) )= ∇ ⋅ v(x, t ) dV (x, t ) ∇⋅v
dt

104
Area differential vector da
 The material time derivative of the area differential associated to a
given particle,

da ( x( X, t ), t ) = F ( X, t ) ⋅ dA( X) ⋅ F −1 ( X, t ) = F ⋅ dA ⋅ F −1

d
dt

dA ⋅ F −1 + F ⋅ dA ( F −1 )
d dF d
(t )
da=
dt dt dt
= F ∇⋅v = −F −1 ⋅ l

d
( da ) = ( ∇ ⋅ v ) F dA ⋅ F −1 − F dA ⋅ F −1 ⋅ l
dt
= da = da

d
( da ) = da ( ∇ ⋅ v ) − da ⋅ l = da ⋅ 1(∇ ⋅ v) − da ⋅ l = da ⋅ ( (∇ ⋅ v)1 − l )
dt da⋅1

105
2.15 Other Coordinate Systems
Ch.2. Deformation and Strain

106
Curvilinear Orthogonal Coord. System
 A curvilinear coordinate system is defined by:
 The coordinates, generically named {a, b, c}
 Its vector basis, {eˆ a , eˆ b , eˆ c } , formed by unit vectors eˆ=
a eˆ=
b e=
ˆ c 1.

 If the elements of the basis are orthogonal is is called an orthogonal


coordinate system: eˆ a ⋅ eˆ b = eˆ a ⋅ eˆ c = eˆ b ⋅ eˆ c = 0
 The orientation of the curvilinear basis may change at each point in
space, .
eˆ m ≡ eˆ m (x) m ∈ {a, b, c}

REMARK
A curvilinear orthogonal coordinate system can be seen as a mobile Cartesian
coordinate system { x′, y′, z ′} , associated to a curvilinear basis {eˆ a , eˆ b , eˆ c } .

108
Curvilinear Orthogonal Coord. System
 A curvilinear orthogonal coordinate system can be seen as a mobile
Cartesian coordinate system {eˆ a , eˆ b , eˆ c } , associated to a curvilinear basis
{ x′, y′, z′} .
 The components of a vector and a tensor magnitude in the curvilinear
orthogonal basis will correspond to those in the given Cartesian local
system:
 va   v x′   Taa Tab Tac  Tx′x′ Tx′y′ Tx′z′ 
     
v ≡  v b  ≡  v y′  T ≡ Tba Tbb Tbc  ≡  Ty′x′ Ty′y′ Ty′z′ 
     Tca Tcc   Tz′x′ Tz′z′ 
 v c   v z′  Tcb Tz′y′

 The components of the curvilinear operators will not be the same as


those in the given Cartesian local system.
 They must be obtained for each specific case.

109
Cylindrical Coordinate System

← z− coordinate line
← θ − coordinate line  x = r cos θ

x( r , θ , z ) ≡  y = r sin θ
← r− coordinate line z = z

∂eˆ r ∂eˆθ
= eˆθ = −eˆ r
∂θ ∂θ

dV = r dθ dr dz

110
Cylindrical Coordinate System
 Nabla operator ∂ 
 ∂r 
 
∂ 1 ∂ ∂  1 ∂ 
∇= eˆ r + eˆθ + eˆ z ⇒ ∇≡
∂r r ∂θ ∂z  r ∂θ   x = r cos θ
  
∂ x( r , θ , z ) ≡  y = r sin θ
 
 ∂z  z = z

 Displacement vector
ur 
u= u r eˆ r + uθ eˆθ + u z eˆ z ⇒ u=  uθ 
 u z 

 Velocity vector  vr 
v= v r eˆ r + vθ eˆθ + v z eˆ z ⇒ u=  vθ 
 v z 

111
Cylindrical Coordinate System
 Infinitesimal strain tensor
ε x′x′ ε x′y′ ε x′z′   ε rr ε rθ ε rz 
ε=
1
2
{
[u ⊗ ∇ ] + [u ⊗ ∇ ]
T
}  
≡ ε x′y′ ε y′y′ ε y′z′ =

ε
 rθ εθθ εθ z 
ε x′z′ ε y′z′ ε z′z′   ε rz εθ z ε zz 
 
∂u 1  1 ∂u r ∂uθ uθ 
ε rr = r ε= + − 
2  r ∂θ

∂r ∂r r 
1 ∂uθ u r 1  ∂u r ∂u z 
εθθ
= + =ε rz  + 
r ∂θ r 2  ∂ z ∂r 
∂u 1  ∂uθ 1 ∂u z 
ε zz = z εθ z
= +
∂z  
2  ∂z r ∂θ 

 x = r cos θ

x( r , θ , z ) ≡  y = r sin θ
z = z

112
Cylindrical Coordinate System
 Strain rate tensor
 d x′x′ d x′y′ d x′z′   d rr d rθ d rz 
d=
1
2
{
[v ⊗ ∇] + [v ⊗ ∇]
T
} 
≡  d x′y′ d y′y′

d y′z′ =
d
 rθ dθθ dθ z 
 d x′z′ d y′z′ d z′z′   d rz dθ z d zz 

∂v 1  1 ∂v r ∂vθ vθ 
d rr = r d= + − 
2  r ∂θ

∂r ∂r r 
1 ∂vθ v r 1  ∂v r ∂v z 
=
dθθ + =
d rz  + 
r ∂θ r 2  ∂ z ∂r 
∂v 1  ∂vθ 1 ∂v z 
d zz = z = +
∂z dθ z  
2  ∂z r ∂θ 

 x = r cos θ

x( r , θ , z ) ≡  y = r sin θ
z = z

113
Spherical Coordinate System

 x = r sin θ cos φ
=x x ( r ,θ , ϕ ) =

≡  y r sin θ sin φ r− coordinate line →
 z = r cos θ

∂eˆ r ∂eˆθ ∂eˆ φ


=
eˆθ =
−eˆ r =
0
∂θ ∂θ ∂θ

dV = r 2 sin θ dr dθ dφ

114
Spherical Coordinate System
 Nabla operator  ∂ 
 ∂ r 
 
∂ 1 ∂ ∂ 1 ∂ 
eˆ φ ⇒ ∇ ≡   x = r sin θ cos φ
1
∇= eˆ r + eˆθ + 
∂r r ∂θ r sin θ ∂φ r ∂θ  = x x ( r , θ , φ ) ≡=
 y r sin θ sin φ
   z = r cos θ
 1 ∂  
 r sin θ ∂φ 

 Displacement vector ur 


 
u= u r eˆ r + uθ eˆθ + uφ eˆ φ ⇒ u=  uθ 
 uφ 
 

 Velocity vector  vr 
 
v= v r eˆ r + vθ eˆθ + vφ eˆ φ ⇒ u=  vθ 
 vφ 
 

115
Spherical Coordinate System
 Infinitesimal strain tensor
ε x′x′ ε x′y′ ε x′z′   ε rr ε rθ ε rφ 
ε=
1
2
{[u ⊗ ∇ ] + [u ⊗ ∇ ]
T
}  
≡ ε x′y′ ε y′y′ ε y′z′ =


εθ r εθθ εθφ 

ε x′z′ ε y′z′ ε z′z′  ε rφ εθφ ε φφ 
  
∂u r
ε rr =
∂r
1 ∂uθ u r
εθθ
= +  x = r sin θ cos φ
r ∂θ r 
1 ∂uφ uθ u =x x ( r , θ , φ ) ≡=
 y r sin θ sin φ
ε ϕϕ= + cotgφ + r  z = r cos θ
r sin θ ∂φ r r 

1  1 ∂u r ∂uθ uθ 
ε= + − 
2  r ∂θ

∂r r 
1  1 ∂u r ∂uφ uφ 
=ε rφ + − 
2  r sin θ ∂φ ∂r r 
1  1 ∂uθ 1 ∂uφ uφ 
=εθφ + − φ
2  r sin θ ∂φ r ∂θ 
cotg
r 

116
Spherical Coordinate System
 Deformation rate tensor
 d x′x′ d x′y′ d x′z′   d rr d rθ d rφ 
d=
1
2
{[v ⊗ ∇] + [v ⊗ ∇]
T
} 
≡  d x′y′ d y′y′

d y′z′ =

 d rθ dθθ dθφ 

 d x′z′ d y′z′ d z′z′   d rφ dθφ dφφ 
 
∂v r
d rr =
∂r  x = r sin θ cos φ

1 ∂vθ v r =x x ( r , θ , φ ) ≡=
 y r sin θ sin φ
=
dθθ +  z = r cos θ
r ∂θ r 
1 ∂vφ vθ v
dφφ = + cotgϕ + r
r sin θ ∂φ r r
1  1 ∂v r ∂vθ vθ 
d= + − 
2  r ∂θ

∂r r 
1  1 ∂v r ∂vφ vφ 
= + − 
2  r sin θ ∂φ
d rφ
∂r r 
1  1 ∂vθ 1 ∂vφ vφ 
= + − φ
2  r sin θ ∂φ r ∂θ 
dθφ cotg
r 

117
Chapter 2
Strain

rs
ee
s gin
2.1 Introduction

t d le En

r
ba
ge ro or
eS m
ci
f
Definition 2.1. In the broader context, the concept of deformation no

ra
C d P cs
longer refers to the study of the absolute motion of the particles as
b
a
i
seen in Chapter 1, but to the study of the relative motion, with respect
an an n

to a given particle, of the particles in its differential neighborhood.


y ha

le
liv or ec
M

.A

2.2 Deformation Gradient Tensor


m

d
uu

Consider the continuous medium in motion of Figure 2.1. A particle P in the


e
X Th


reference configuration Ω0 occupies the point in space P in the present config-
er
tin

uration Ωt , and a particle Q situated in the differential neighborhood of P has


on

.O

relative positions with respect to this particle in the reference and present times
given by dX and dx, respectively. The equation of motion is given by
C


©

not
x = ϕ (X,t) = x (X,t)
not . (2.1)
xi = ϕi (X1 , X2 , X3 ,t) = xi (X1 , X2 , X3 ,t) i ∈ {1, 2, 3}

Differentiating (2.1) with respect to the material coordinates X results in the



Fundamental ⎪
⎨ dx = F · dX
equation of
⎪ ∂ xi (2.2)
deformation ⎩ dxi = dX j = Fi j dX j i, j ∈ {1, 2, 3}
∂ Xj

41
42 C HAPTER 2. S TRAIN

rs
Figure 2.1: Continuous medium in motion.

ee
s gin
Equation (2.2) defines the material deformation gradient tensor F (X,t) 1 .

t d le En

⎪ not
⎨F = x⊗∇

r
ba
Material deformation

ge ro or
∂ xi

eS m
(2.3)
gradient tensor ⎪
⎩ Fi j = i, j ∈ {1, 2, 3}

ci
∂ Xj f

ra
C d P cs
b
a
i
an an n

The explicit components of tensor F are given by


y ha

⎡ ⎤
∂ x1 ∂ x1 ∂ x1
le
liv or ec

⎡ ⎤ ⎢ ⎥
x  ⎢ ∂X ∂ X2 ∂ X3 ⎥
 ⎢ 1⎥  ∂
M

⎢ 1 ⎥
.A

 ∂ ∂ ⎢ ∂ x2 ∂ x2 ∂ x2 ⎥
[F] = x ⊗ ∇ = ⎢ ⎥
⎣ x2 ⎦ ∂ X1 , ∂ X2 , ∂ X3 = ⎢ ⎥ . (2.4)
⎢ ∂ X1 ∂ X3 ⎥
m

∂ X2
 ⎢ ⎥
d

x3   ⎣ ∂x ∂ x3 ⎦
uu

 T ∂ x3
e

   3
X Th


er

∂ ∂ X2 ∂ X3
tin

[x] X1
on

.O
C

Remark 2.1. The deformation gradient tensor F (X,t) contains the


©

information of the relative motion, along time t, of all the material


particles in the differential neighborhood of a given particle, identi-
fied by its material coordinates X. In effect, equation (2.2) provides
the evolution of the relative position vector dx in terms of the cor-
responding relative position in the reference time, dX. Thus, if the
value of F (X,t) is known, the information associated with the gen-
eral concept of deformation defined in Section 2.1 is also known.

1 Here, the symbolic form of the material Nabla operator, ∇ ≡ ∂ êi /∂ Xi , applied to the
not
expression of the open or tensor product, [a ⊗ b]i j = [a b]i j = ai b j , is considered.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Deformation Gradient Tensor 43

2.2.1 Inverse Deformation Gradient Tensor


Consider now the inverse equation of motion
 not
X = ϕ −1 (x,t) = X (x,t) ,
(2.5)
Xi = ϕi−1 (x1 , x2 , x3 ,t) = Xi (x1 , x2 , x3 ,t)
not
i ∈ {1, 2, 3} .

Differentiating (2.5) with respect to the spatial coordinates xi results in




⎨ dX = F−1 · dx ,

rs
⎪ ∂ Xi (2.6)
⎩ dXi = dx j = Fi−1
j dx j i, j ∈ {1, 2, 3} .

ee
∂xj

s gin
The tensor defined in (2.6) is named spatial deformation gradient tensor or in-
verse (material) deformation gradient tensor and is characterized by2

t d le En

r
ba

ge ro or
not
⎨ F−1 = X ⊗ ∇

eS m
ci
Spatial deformation
∂ Xi f (2.7)

ra
gradient tensor ⎪ −1
⎩ Fi j = i, j ∈ {1, 2, 3}
C d P cs
∂xj
b
a
i
an an n
y ha

le
liv or ec

Remark 2.2. The spatial deformation gradient tensor, denoted in


M

(2.6) and (2.7) as F−1 , is in effect the inverse of the (material) defor-
.A

mation gradient tensor F. The verification is immediate since3


m

∂ xi ∂ Xk ∂ xi not
uu
e

= = δi j =⇒ F · F−1 = 1 ,
X Th

∂ Xk ∂ x j ∂xj
er
tin

 
Fik F −1
on

.O

kj
C

∂ Xi ∂ xk ∂ Xi not
F−1 · F = 1 .
©

= = δi j =⇒
∂ xk ∂ X j ∂ Xj
 
Fik−1 Fk j

2 Here, the symbolic form of the spatial Nabla operator, ∇ ≡ ∂ êi /∂ xi , is considered. Note
the difference in notation between this spatial operator ∇ and the material Nabla ∇.
3 The two-index operator Delta Kronecker δ is defined as δ = 1 if i = j and δ = 0 if
ij ij ij
i = j. The second-order unit tensor 1 is given by [1]i j = δi j .

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
44 C HAPTER 2. S TRAIN

The explicit components of tensor F−1 are given by


⎡ ⎤
∂ X1 ∂ X1 ∂ X1
⎡ ⎤ ⎢ ⎥
X1   ⎢ ∂x ∂ x2 ∂ x3 ⎥
 −1  ⎢ ⎥ ∂ ∂ ∂ ⎢ 1 ⎥
⎢ ∂ X2 ∂ X2 ∂ X2 ⎥
F = [X ⊗ ∇] = ⎢ ⎥
⎣ X2 ⎦ ∂ x1 , ∂ x2 , ∂ x3 = ⎢ ⎥ . (2.8)
⎢ ∂ x1 ∂ x2 ∂ x3 ⎥
X3    ⎢⎣ ∂X

∂ X3 ∂X ⎦
   [∇]T
3 3

[X] ∂ x1 ∂ x2 ∂ x3

rs
ee
Example 2.1 – At a given time, the motion of a continuous medium is defined

s gin
by ⎧
⎨ x1 = X1 − AX3

t d le En
x2 = X2 − AX3 .

r

ba
x3 = −AX1 + AX2 + X3

ge ro or
eS m
ci
f
Obtain the material deformation gradient tensor F (X,t) at this time. By

ra
C d P cs
b
a
means of the inverse equation of motion, obtain the spatial deformation gra-
i
dient tensor F−1 (x). Using the results obtained, verify that F · F−1 = 1.
an an n
y ha

le
Solution
liv or ec
M

.A

The material deformation gradient tensor is


⎡ ⎤
m

X1 − AX3
d

 T ⎢  
⎥ ∂
uu

∂ ∂
e

not
F = x ⊗ ∇ ≡ [x] ∇ = ⎣ ⎢ ⎥
X2 − AX3 ⎦ ∂ X1 , ∂ X2 , ∂ X3
X Th

er
tin

−AX1 + AX2 + X3
on

.O

⎡ ⎤
−A
C

1 0
not ⎢ ⎥
©

F≡⎣ 0 1 −A ⎦ .
−A A 1
The inverse equation of motion is obtained directly from the algebraic inver-
sion of the equation of motion,
⎡   ⎤
X1 = 1 + A2 x1 − A2 x2 + Ax3
not ⎢   ⎥
X (x,t) ≡ ⎢ ⎥
⎣ X2 = A x1 + 1 − A x2 + Ax3 ⎦ .
2 2

X3 = Ax1 − Ax2 + x3

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Displacements 45

Then, the spatial deformation gradient tensor is


⎡  ⎤
1 + A2 x1 − A2 x2 + Ax3  
⎢   ⎥ ∂ ∂ ∂
−1 not T ⎢
F = X ⊗ ∇ ≡ [X] [∇] = ⎣ A x1 + 1 − A x2 + Ax3 ⎦
2 2 ⎥ , ,
∂ x1 ∂ x2 ∂ x3
Ax1 − Ax2 + x3
⎡ ⎤
1 + A2 −A2 A
not ⎢ ⎥
F−1 ≡ ⎣ A2 1 − A2 A⎦ .

rs
A −A 1

ee
Finally, it is verified that

s gin
⎡ ⎤⎡ ⎤ ⎡ ⎤
0 −A 1 + A2 −A2

t d le En
1 A 1 0 0
not ⎢ ⎥⎢ ⎥ ⎢ ⎥ not
F · F−1 ≡ ⎣ 0 1 −A ⎦ ⎣ A2 1 − A2 A⎦ = ⎣0 1 0⎦ ≡ 1 .

r
ba
ge ro or
eS m
−A A 1 A −A 1 0 0 1

ci
f

ra
C d P cs
b
a
i
an an n
y ha

2.3 Displacements
le
liv or ec
M

.A

Definition 2.2. A displacement is the difference between the posi-


m

tion vectors in the present and reference configurations of a same


d
uu

particle.
e
X Th

er
tin
on

.O

The displacement of a particle P at a given time is defined by vector u, which



C

joins the points in space P (initial position) and P (position at the present time t)
©

of the particle (see Figure 2.2). The displacement of all the particles in the con-
tinuous medium defines a displacement vector field which, as all properties of
the continuous medium, can be described in material form U (X,t) or in spatial
form u (x,t) as follows.

U (X,t) = x (X,t) − X
(2.9)
Ui (X,t) = xi (X,t) − Xi i ∈ {1, 2, 3}

u (x,t) = x − X (x,t)
(2.10)
ui (x,t) = xi − Xi (x,t) i ∈ {1, 2, 3}

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
46 C HAPTER 2. S TRAIN

rs
ee
s gin
Figure 2.2: Displacement of a particle.

t d le En

r
2.3.1 Material and Spatial Displacement Gradient Tensors

ba
ge ro or
eS m
ci
Differentiation with respect to the material coordinates of the displacement vec-
f

ra
C d P cs
tor Ui defined in (2.9) results in
b
a
i
∂Ui ∂ xi ∂ Xi
an an n

de f
= − = Fi j − δi j = Ji j , (2.11)
y ha

∂ Xj ∂ Xj ∂ Xj
 
le
liv or ec

Fi j δi j
M

.A

which defines the material displacement gradient tensor as follows.


m


uu


e

de f
⎨ J (X,t) = U (X,t) ⊗ ∇ = F − 1
X Th

er

Material displacement
tin

gradient tensor ⎪ ∂Ui (2.12)


⎩ Ji j = = Fi j − δi j i, j ∈ {1, 2, 3}
∂ Xj
on

.O
C



©

⎨ U = J · dX
⎪ ∂Ui (2.13)
⎩ dUi = dX j = Ji j dX j i, j ∈ {1, 2, 3}
∂ Xj
Similarly, differentiation with respect to the spatial coordinates of the expres-
sion of ui given in (2.10) yields
∂ ui ∂ xi ∂ Xi de f
= − = δi j − Fi−1
j = ji j , (2.14)
∂xj ∂xj ∂xj
 
δi j Fi−1
j

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Strain Tensors 47

which defines the spatial displacement gradient tensor as follows.



⎪ de f
⎨ j (x,t) = u (x,t) ⊗ ∇ = 1 − F−1
Spatial displacement
⎪ ∂ ui (2.15)
gradient tensor ⎩ ji j = = δi j − Fi−1 i, j ∈ {1, 2, 3}
∂xj j



⎨ u = j · dx
⎪ ∂ ui (2.16)
⎩ dui = dx j = ji j dx j i, j ∈ {1, 2, 3}

rs
∂xj

ee
s gin
2.4 Strain Tensors

t d le En
Consider now a particle of the continuous medium that occupies the point in

r
space P in the material configuration, and another particle Q√in its differen-

ba
ge ro or
eS m
dS = dX · dX) from

ci
tial neighborhood separated a segment dX (with length

f

ra
the previous paticle, being dx (with length ds = dx · dx) its counterpart in
C d P cs
b
a
the present configuration (see Figure 2.3). Both differential vectors are related
i
an an n

through the deformation gradient tensor F (X,t) by means of equations (2.2) and
y ha

(2.6),
le

liv or ec

⎨ dx = F · dX and dX = F−1 · dx ,
M

.A

(2.17)
⎩ dxi = Fi j dX j and dXi = F −1 dx j i, j ∈ {1, 2, 3} .
m

ij
d
uu
e

Then,
X Th


er
tin

⎨ (ds)2 = dx · dx not
≡ [dx]T [dx] = [F · dX]T [F · dX] ≡ dX · FT · F · dX
not
on

.O

(2.18)
⎩ (ds)2 = dxk dxk = Fki dXi Fk j dX j = dXi Fki Fk j dX j = dXi F T Fk j dX j
ik
C

or, alternatively4 ,
⎧  −1 T  −1 

⎪ (dS) 2
= dX · dX
not
≡ [dX] T
[dX] = F · dx F · dx =



⎨ ≡ dx · F−T · F−1 · dx ,
not

(2.19)

⎪ (dS)2 = dXk dXk = Fki−1 dxi Fk−1 −1 −1

⎪ j dx j = dxi Fki Fk j dx j =


= dxi Fik−T Fk−1
j dx j .

 T not
4 The convention (•)−1 = (•)−T is used.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
48 C HAPTER 2. S TRAIN

rs
ee
s gin
Figure 2.3: Differential segments in a continuous medium.

t d le En

r
2.4.1 Material Strain Tensor (Green-Lagrange Strain Tensor)

ba
ge ro or
eS m
ci
Subtracting expressions (2.18) and (2.19) results in
f

ra
C d P cs
b
a
(ds)2 − (dS)2 = dX · FT · F · dX − dX · dX =
i
an an n

= dX · FT · F · dX − dX · 1 · dX =
y ha

 
le
= dX · FT · F − 1 · dX = 2 dX · E · dX , (2.20)
liv or ec

  
M

.A

de f
= 2E
m

which implicitly defines the material strain tensor or Green-Lagrange strain


uu

tensor as follows.
e
X Th

er


tin

⎪  
Material ⎨ E (X,t) = 1 FT · F − 1
on

.O

2 (2.21)
(Green-Lagrange)
⎪ 1 
strain tensor ⎩ Ei j (X,t) = Fki Fk j − δi j i, j ∈ {1, 2, 3}
C

2
©

Remark 2.3. The material strain tensor E is symmetric. Proof is ob-


tained directly from (2.21), observing that
⎧  
⎨ ET = 1 FT · F − 1T = 1 FT · FT T − 1T = 1 FT · F − 1 = E ,
2 2 2
⎩E = E i, j ∈ {1, 2, 3} .
ij ji

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Strain Tensors 49

2.4.2 Spatial Strain Tensor (Almansi Strain Tensor)


Subtracting expressions (2.18) and (2.19) in an alternative form yields

(ds)2 − (dS)2 = dx · dx − dx · F−T · F−1 · dx =


= dx · 1 · dx − dx · F−T · F−1 · dx =
 
= dx · 1 − F−T · F−1 · dx = 2 dx · e · dx , (2.22)
  
de f
= 2e

rs
which implicitly defines the spatial strain tensor or Almansi strain tensor as

ee
follows.

s gin
⎪  
Spatial ⎨ e (x,t) = 1 1 − F−T · F−1

t d le En
(Almansi) 2   (2.23)

strain tensor ⎩ ei j (x,t) =
1
δi j − Fki−1 Fk−1 i, j ∈ {1, 2, 3}

r
2 j

ba
ge ro or
eS m
ci
f

ra
C d P cs
b
a
i
Remark 2.4. The spatial strain tensor e is symmetric. Proof is ob-
an an n

tained directly from (2.23), observing that


y ha

⎧ 
⎪     
T = 1 1 − F−T · F−1 T = 1 1T − F−1 T · F−T T =
 
le

liv or ec


⎪ e
⎨ 2 2
M

.A

1 −T −1


⎪ = 1−F ·F =e,
⎪ 2
m


⎩e = e
d

ij ji i, j ∈ {1, 2, 3} .
uu
e
X Th

er
tin
on

.O

Example 2.2 – Obtain the material and spatial strain tensors for the motion
C

in Example 2.1.
©

Solution
The material strain tensor is ⎛⎡ ⎤⎡ ⎤ ⎡ ⎤⎞
1 0 −A 1 0 −A 1 0 0
1 T  not 1
E (X,t) = F · F − 1 ≡ ⎝⎣ 0 1 A ⎦⎣ 0 1 −A ⎦−⎣ 0 1 0 ⎦⎠ =
2 2
−A −A 1 −A A 1 001
⎡ 2 ⎤
A −A2 −2A
1
= ⎣ −A2 A2 0 ⎦
2
−2A 0 2A2

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
50 C HAPTER 2. S TRAIN

and the spatial strain tensor is


1 
e (X,t) = 1 − F−T · F−1 =
2
⎛⎡ ⎤ ⎡ ⎤⎡ ⎤⎞
1 0 0 1 + A2 A2 A 1 + A2 −A2 A
not 1
≡ ⎝⎣ 0 1 0 ⎦ − ⎣ −A2 1 − A2 −A ⎦ ⎣ A2 1 − A2 A ⎦⎠ =
2
0 0 1 A A 1 A −A 1
⎡ ⎤
−3A2 − 2A4 A2 + 2A4 −2A − 2A3
1
= ⎣ A2 + 2A4 A2 − 2A4 ⎦.

rs
2A3
2
−2A − 2A −2A

ee
3 2A3 2

s gin
Observe that E = e.

t d le En

r
ba
ge ro or
eS m
ci
f

ra
C d P cs
Remark 2.5. The material strain tensor E and the spatial strain ten-
b
a
sor e are different tensors. They are not the material and spatial de-
i
an an n

scriptions of a same strain tensor. Expressions (2.20) and (2.22),


y ha

le
(ds)2 − (dS)2 = 2dX · E · dX = 2dx · e · dx ,
liv or ec
M

.A

clearly show this since each tensor is affected by a different vector


(dX and dx, respectively).
m

d
uu

The Green-Lagrange strain tensor is naturally described in mate-


e

rial description (E (X,t)). In equation (2.20) it acts on element dX


X Th

er
tin

(defined in material configuration) and, hence, its denomination as


material strain tensor. However, as all properties of the continuous
on

.O

medium, it may be described, if required, in spatial form (E (x,t))


C

through the adequate substitution of the equation of motion.


©

The contrary occurs with the Almansi strain tensor: it is naturally


described in spatial form and in equation (2.22) acts on the differ-
ential vector dx (defined in the spatial configuration) and, thus, its
denomination as spatial strain tensor. It may also be described, if
required, in material form (e (X,t)).

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Variation of Distances: Stretch and Unit Elongation 51

2.4.3 Strain Tensors in terms of the Displacement (Gradients)


Replacing expressions (2.12) and (2.15) into equations (2.21) and (2.23) yields
the expressions of the strain tensors in terms of the material displacement gradi-
ent, J (X,t), and the spatial displacement gradient, j (x,t).

1   1 
E= 1 + JT · (1 + J) − 1 = J + JT + JT · J
2 2
 (2.24)
1 ∂Ui ∂U j ∂Uk ∂Uk
Ei j = + + i, j ∈ {1, 2, 3}
2 ∂ X j ∂ Xi ∂ Xi ∂ X j

rs
ee
1  1

s gin
  
e= 1 − 1 − jT · (1 − j) = j + jT − jT · j
2 2


t d le En
(2.25)
1 ∂ ui ∂ u j ∂ uk ∂ uk
ei j = + − i, j ∈ {1, 2, 3}

r
2 ∂ x j ∂ xi ∂ xi ∂ x j

ba
ge ro or
eS m
ci
f

ra
C d P cs
b
a
i
an an n

2.5 Variation of Distances: Stretch and Unit Elongation


y ha

Consider now a particle P in the reference configuration and another particle


le
liv or ec

Q, belonging to the differential neighborhood of P (see Figure 2.4). The corre-


M

.A

sponding positions in the present configuration are given by the points in space
 
P and Q such that the distance between the two particles in the reference con-
m

figuration, dS, is transformed into ds at the present time. The vectors T and t are
uu

 
e

the unit vectors in the directions PQ and P Q , respectively.


X Th

er
tin
on

.O

Definition 2.3. The stretch or stretch ratio of a material point P (or



C

a spatial point P ) in the material direction T (or spatial direction t )


©

 
is the length of the deformed differential segment P Q per unit of
length of the original differential segment PQ.

The translation of the previous definition into mathematical language is

 
de f PQ ds
Stretch = λT = λt = = (0 < λ < ∞) . (2.26)
PQ dS

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
52 C HAPTER 2. S TRAIN

rs
ee
s gin
Figure 2.4: Differential segments and unit vectors in a continuous medium.

t d le En

r
ba
ge ro or
eS m
ci
f
Definition 2.4. The unit elongation, elongation ratio or extension of

ra

C d P cs
a material point P (or a spatial point P ) in the material direction T
b
a
i
5
(or spatial direction t ) is the increment of length of the deformed
an an n

 
y ha

differential segment P Q per unit of length of the original differen-


le
tial segment PQ.
liv or ec
M

.A
m

The corresponding mathematical definition is


d
uu
e
X Th

Δ PQ ds − dS
er

de f
tin

Unit elongation = εT = εt = = . (2.27)


PQ dS
on

.O
C

Equations (2.26) and (2.27) allow immediately relating the values of the unit
©

elongation and the stretch for a same point and direction as follows.
ds − dS ds
ε= = −1 = λ − 1 (⇒ −1 < ε < ∞) (2.28)
dS dS

λ

5 Often, the subindices (•)T and (•)t will be dropped when referring to stretches or unit
elongations. However, one must bear in mind that both stretches and unit elongations are
always associated with a particular direction.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Variation of Distances: Stretch and Unit Elongation 53

Remark 2.6. The following deformations may take place:


• If λ = 1 (ε = 0) ⇒ ds = dS: The particles P and Q may have
moved along time, but without increasing or decreasing the dis-
tance between them.
• If λ > 1 (ε > 0) ⇒ ds > dS: The distance between the particles
P and Q has lengthened with the deformation of the medium.
• If λ < 1 (ε < 0) ⇒ ds < dS: The distance between the particles

rs
P and Q has shortened with the deformation of the medium.

ee
s gin
t d le En
2.5.1 Stretches, Unit Elongations and Strain Tensors

r
ba
ge ro or
Consider equations (2.21) and (2.22) as well as the geometric expressions

eS m
ci
dX = T dS and dx = t ds (see Figure 2.4). Then,
f

ra
C d P cs

b
a

⎪ (ds)2 − (dS)2 = 2 dX · E · dX = 2 (dS)2 T · E · T
i

an an n


⎨  
y ha

dS T dS T (2.29)
le

⎪ (ds)2 − (dS)2 = 2 dx · e · dx = 2 (ds)2 t · e · t
liv or ec


⎪  

M

.A

ds t ds t
m

and dividing these expressions by (dS)2 and (ds)2 , respectively, results in


d
uu
e


X Th


er
tin

ds 2 λ = 1 + 2T · E · T
−1 = λ −1 = 2 T·E·T ⇒
2
√ (2.30)
dS ε = λ − 1 = 1 + 2T · E · T − 1
on

.O

  
C

λ
©

  1
λ=√
dS 2 1 2
1 − 2t · e · t
1− = 1− = 2 t·e·t ⇒ 1 (2.31)
ds λ ε = λ −1 = √ −1
   1 − 2t · e · t
1/λ

These equations allow calculating the unit elongation and stretch for a given
direction (in material description, T, or in spatial description, t ).

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
54 C HAPTER 2. S TRAIN

Remark 2.7. The material and spatial strain tensors, E (X,t) and
e (x,t), contain information on the stretches (and unit elongations)
for any direction in a differential neighborhood of a given particle,
as evidenced by (2.30) and (2.31).

Example 2.3 – The spatial strain tensor for a given motion is

rs
⎡ ⎤

ee
0 0 −tetz
not ⎢ ⎥

s gin
e (x,t) ≡ ⎣ 0 0 0 ⎦.
−tetz t (2etz − et )

t d le En
0

r
Calculate the length, at time t = 0, of the segment that at time t = 2 is recti-

ba
ge ro or
eS m
linear and joins points a ≡ (0, 0, 0) and b ≡ (1, 1, 1).

ci
f

ra
C d P cs
b
a
Solution
i
an an n

The shape and geometric position of the material segment at time t = 2 is


y ha

known. At time t = 0 (reference time) the segment is not necessarily recti-


le
liv or ec

linear and the positions of its extremes A and B (see figure below) are not
M

.A

known. To determine its length, (2.31) is applied for a unit vector in the di-
rection of the spatial configuration t,
m

1 ds 1
uu

λ=√ = =⇒ dS = ds .
e

λ
X Th

1 − 2 t · e · t dS
er
tin
on

.O
C

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Variation of Angles 55


To obtain the stretch in the direction t ≡ [1, 1, 1]T / 3, the expression t · e · t
not

is computed first as
⎡ ⎤⎡ ⎤
0 0 −tetz 1
not 1 ⎢ ⎥⎢ ⎥ 1 1 t
t·e·t ≡ √ [1, 1, 1] ⎣ 0 0 0 ⎦ ⎣ 1 ⎦ √ = − te .
3 3 3
−tetz 0 t (2etz − et ) 1

Then, the corresponding stretch at time t = 2 is


" √

rs
1 " 1 3
λ=! =⇒ λ" =! =√ .

ee
1 + 23 tet t=2
1+ 3e
4 2 3 + 4e2

s gin
The length at time t = 0 of the segment AB is

t d le En
# B # b # b
1 1 1 1√

r
lAB = dS = ds = ds = lab =

ba
3

ge ro or
λ λ λ λ

eS m
A a a
  

ci
f

ra
lab
C d P cs
b
a
i
and replacing the expression obtained above for the stretch at time t = 2
an an n
y ha

finally results in $
le
lAB = 3 + 4e2 .
liv or ec
M

.A
m

d
uu

2.6 Variation of Angles


e
X Th

er
tin

Consider a particle P and two additional particles Q and R, belonging to the dif-
ferential neighborhood of P in the material configuration (see Figure 2.5), and
on

.O

  
the same particles occupying the spatial positions P , Q and R . The relationship
C

between the angles that form the corresponding differential segments in the ref-
©

erence configuration (angle Θ ) and the present configuration (angle θ ) is to be


considered next.
Applying (2.2) and (2.6) on the differential vectors that separate the particles,
⎧ ⎧
⎨ dx(1) = F · dX(1) ⎨ dX(1) = F−1 · dx(1)
=⇒ (2.32)
⎩ dx(2) = F · dX(2) ⎩ dX(2) = F−1 · dx(2)

and using the definitions of the unit vectors T(1) , T(2) , t(1) and t(2) that establish
the corresponding directions in Figure 2.5,

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
56 C HAPTER 2. S TRAIN

rs
ee
s gin
t d le En

r
ba
ge ro or
eS m
Figure 2.5: Angles between particles in a continuous medium.

ci
f

ra
C d P cs
⎧ ⎧
b
a
⎨ dX(1) = dS(1) T(1) ⎨ dx(1) = ds(1) t(1) ,
i
an an n

=⇒ (2.33)
y ha

⎩ dX(2) = dS(2) T(2) ⎩ dx(2) = ds(2) t(2) .


le
liv or ec
M

Finally, according to the definition in (2.26), the corresponding stretches are


.A



m


⎪ 1
d

⎨ ds(1) = λ (1) dS(1) ⎨ dS(1) = ds(1) ,


uu

λ (1)
e

=⇒ (2.34)
X Th

⎩ ds(2) = λ (2) dS(2) ⎪


er


tin

⎩ dS(2) = 1
ds(2) .
λ (2)
on

.O

Expanding now the scalar product6 of the vectors dx(1) and dx(2) ,
C

" "" "  T  


©

" "" "


dS(1) dS(2) cos θ = "dx(1) " "dx(2) " cos θ = dx(1) · dx(2) ≡ dx(1) dx(2) =
not

 T    
= F · dX(1) F · dX(2) ≡ dX(1) · FT · F · dX(2) = dX(1) · (2E + 1) · dX(2)
not

1 1
= dS(1) T(1) · (2E + 1) · T(2) dS(2) = (1) ds(1) T(1) · (2E + 1) · T(2) (2) ds(2) =
λ λ
(1) (2) 1 1 (1) (2)
= ds ds T · (2E + 1) · T ,
λ (1) λ (2)
(2.35)

6 The scalar product of two vectors a and b is defined in terms of the angle between them, θ ,
as a · b = |a| · |b| cos θ .

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Physical Interpretation of the Strain Tensors 57

and, comparing the initial and final terms in (2.35), yields

T(1) · (1 + 2E) · T(2)


cos θ = , (2.36)
λ (1) λ (2)
where the stretches λ (1) and λ (2) can be obtained by applying (2.30) to the
directions T(1) and T(2) , resulting in

T(1) · (1 + 2E) · T(2)


cos θ = $ $ . (2.37)
1 + 2T(1) · E · T(1) 1 + 2T(2) · E · T(2)

rs
In an analogous way, operating on the reference configuration, the angle Θ

ee
between the differential segments dX(1) and dX(2) (in terms of t(1) , t(2) and e )

s gin
is obtained,

t d le En
t(1) · (1 − 2e) · t(2)
cosΘ = $ $ .

r
(2.38)

ba
ge ro or
1 − 2t(1) · e · t(1) 1 − 2t(2) · e · t(2)

eS m
ci
f

ra
C d P cs
b
a
i
Remark 2.8. Similarly to the discussion in Remark 2.7, the material
an an n

and spatial strain tensors, E (X,t) and e (x,t), also contain informa-
y ha

tion on the variation of the angles between differential segments in


le
liv or ec

the differential neighborhood of a particle during the deformation


M

process. These facts will be the basis for providing a physical inter-
.A

pretation of the components of the strain tensors in Section 2.7.


m

d
uu
e
X Th

er
tin

2.7 Physical Interpretation of the Strain Tensors


on

.O
C

2.7.1 Material Strain Tensor


©

Consider a segment PQ, oriented parallel to the X1 -axis in the reference config-
uration (see Figure 2.6). Before the deformation takes place, PQ has a known
length dS = dX.
 
The length of P Q is sought. To this aim, consider the material strain tensor
E given by its components,
⎡ ⎤ ⎡ ⎤
EXX EXY EXZ E11 E12 E13
not ⎢ ⎥ ⎢ ⎥
E ≡ ⎣ EXY EYY EY Z ⎦ = ⎣ E12 E22 E23 ⎦ . (2.39)
EXZ EY Z EZZ E13 E23 E33

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
58 C HAPTER 2. S TRAIN

rs
ee
Figure 2.6: Differential segment in the reference configuration.

s gin
t d le En
Consequently,

r
ba
⎡ ⎤⎡ ⎤

ge ro or
eS m
E11 E12 E13 1

ci
⎣ f⎦ ⎣ 0 ⎦ = E11 .

ra
not T
T · E · T ≡ [T] [E] [T] = [1, 0, 0] E12 E22 E23 (2.40)
C d P cs
b
a
E13 E23 E33 0
i
an an n
y ha

The stretch in the material direction X1 is now obtained by replacing


√ the value
le
T · E · T into the expression for stretch (2.30), resulting in λ1 = 1 + 2E11 . In an
liv or ec

analogous manner, the segments oriented in the directions X2 ≡ Y and X3 ≡ Z


M

.A

are considered to obtain the values λ2 and λ3 as follows.


m

√ √ √
d

λ1 = 1 + 2E11 = 1 + 2EXX ⇒ εX = λX − 1 = 1 + 2EXX − 1


uu
e

√ √ √
X Th

er

λ2 = 1 + 2E22 = 1 + 2EYY ⇒ εY = λY − 1 = 1 + 2EYY − 1


tin

(2.41)
√ √ √
λ3 = 1 + 2E33 = 1 + 2EZZ ⇒ εZ = λZ − 1 = 1 + 2EZZ − 1
on

.O
C

Remark 2.9. The components EXX , EYY and EZZ (or E11 , E22 and
E33 ) of the main diagonal of tensor E (denoted longitudinal strains)
contain the information on stretch and unit elongations of the dif-
ferential segments that were initially (in the reference configuration)
oriented in the directions X, Y and Z, respectively.
• If EXX = 0 ⇒ εX = 0 : No unit elongation in direction X.
• If EYY = 0 ⇒ εY = 0 : No unit elongation in direction Y .
• If EZZ = 0 ⇒ εZ = 0 : No unit elongation in direction Z.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Physical Interpretation of the Strain Tensors 59

rs
ee
Figure 2.7: Angles between differential segments in the reference and present configu-

s gin
rations.

t d le En

r
Consider now the angle between segments PQ (parallel to the X1 -axis) and PR

ba
ge ro or
eS m
(parallel to the X2 -axis), where Q and R are two particles in the differential neigh-

ci
f  

ra
borhood of P in the material configuration and P , Q and R are the respective
C d P cs
b
positions in the spatial configuration (see Figure 2.7). If the angle (Θ = π/2)

a
i
between the segments in the reference configuration is known, the angle θ in
an an n
y ha

the present configuration can be determined using (2.37) and taking into ac-
count their orthogonality ( T(1) · T(2) = 0 ) and the equalities T(1) · E · T(1) = E11 ,
le
liv or ec

T(2) · E · T(2) = E22 and T(1) · E · T(2) = E12 . That is,


M

.A

T(1) · (1 + 2E) · T(2)


m

cos θ = $ $
d
uu

1 + 2T(1) · E · T(1) 1 + 2T(2) · E · T(2)


e

(2.42)
X Th

er
tin

2E12
=√ √ ,
1 + 2E11 1 + 2E22
on

.O

which is the same as


C

π 2EXY
θ ≡ θxy = − arcsin √ √ . (2.43)
2 1 + 2EXX 1 + 2EYY
The increment of the final angle with respect to its initial value results in
2EXY
ΔΘXY = θxy − ΘXY = − arcsin √ √ . (2.44)
 1 + 2EXX 1 + 2EYY
π/2

Analogous results are obtained starting from pairs of segments that are ori-
ented in different combinations of the coordinate axes, resulting in

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
60 C HAPTER 2. S TRAIN

2EXY
ΔΘXY = − arcsin √ √
1 + 2EXX 1 + 2EYY
2EXZ
ΔΘXZ = − arcsin √ √ . (2.45)
1 + 2EXX 1 + 2EZZ
2EY Z
ΔΘY Z = − arcsin √ √
1 + 2EYY 1 + 2EZZ

rs
ee
Remark 2.10. The components EXY , EXZ and EY Z (or E12 , E13 and

s gin
E23 ) of the tensor E (denoted angular strains) contain the informa-
tion on variation of the angles between the differential segments that

t d le En
were initially (in the reference configuration) oriented in the direc-
tions X, Y and Z, respectively.

r
ba
ge ro or
eS m
• If EXY = 0 : The deformation does not produce a variation in the

ci
f
angle between the two segments initially oriented in the direc-

ra
C d P cs
tions X and Y .
b
a
i
• If EXZ = 0 : The deformation does not produce a variation in the
an an n
y ha

angle between the two segments initially oriented in the direc-


tions X and Z.
le
liv or ec

• If EY Z = 0 : The deformation does not produce a variation in the


M

.A

angle between the two segments initially oriented in the direc-


tions Y and Z.
m

d
uu
e
X Th

er
tin

The physical interpretation of the components of the material strain tensor is


shown in Figure 2.8 on an elemental parallelepiped in the neighborhood of a
on

.O

particle P with edges oriented in the direction of the coordinate axes.


C

2.7.2 Spatial Strain Tensor


Arguments similar to those of the previous subsection allow interpreting the
spatial components of the strain tensor,
⎡ ⎤ ⎡ ⎤
exx exy exz e11 e12 e13
not ⎢ ⎥ ⎢ ⎥
e ≡ ⎣ exy eyy eyz ⎦ = ⎣ e12 e22 e23 ⎦ . (2.46)
exz eyz ezz e13 e23 e33

The components of the main diagonal (longitudinal strains) can be interpreted


in terms of the stretches and unit elongations of the differential segments ori-

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Physical Interpretation of the Strain Tensors 61

rs
ee
s gin
Figure 2.8: Physical interpretation of the material strain tensor.

t d le En

r
ented in the direction of the coordinate axes in the present configuration,

ba
ge ro or
eS m
ci
f

ra
C d P cs
1 1 1
b
a
λ1 = √ =√ ⇒ εx = √ −1
i
1 − 2e11 1 − 2exx 1 − 2exx
an an n
y ha

1 1 1
λ2 = √ =$ ⇒ εy = $ −1 ,
le
(2.47)
liv or ec

1 − 2e22 1 − 2eyy 1 − 2eyy


M

.A

1 1 1
λ3 = √ =√ ⇒ εz = √ −1
m

1 − 2e33 1 − 2ezz 1 − 2ezz


d
uu
e
X Th

er

while the components outside the main diagonal (angular strains) contain infor-
tin

mation on the variation of the angles between the differential segments oriented
on

.O

in the direction of the coordinate axes in the present configuration,


C

π 2exy
Δ θxy = −ΘXY = − arcsin √ $
2 1 − 2exx 1 − 2eyy
π 2exz .
Δ θxz = −ΘXZ = − arcsin √ √ (2.48)
2 1 − 2exx 1 − 2ezz
π 2eyz
Δ θyz = −ΘY Z = − arcsin $ √
2 1 − 2eyy 1 − 2ezz

Figure 2.9 summarizes the physical interpretation of the components of the spa-
tial strain tensor.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
62 C HAPTER 2. S TRAIN

rs
Figure 2.9: Physical interpretation of the spatial strain tensor.

ee
s gin
2.8 Polar Decomposition

t d le En
The polar decomposition theorem of tensor analysis establishes that, given a

r
ba
second-order tensor F such that |F| > 0, there exist an orthogonal tensor Q 7 and

ge ro or
eS m
ci
two symmetric tensors U and V such that8
f

ra
C d P cs
not √

b
a
U = FT · F ⎪

i

an an n

not √
y ha

V = F·F T =⇒ F = Q · U = V · Q . (2.49)


le

liv or ec

Q = F · U−1 = V−1 · F
M

.A

This decomposition is unique for each tensor F and is denominated left polar
m

decomposition (F = Q · U) or right polar decomposition (F = V · Q). Tensors U


d
uu

and V are named right and left stretch tensors, respectively.


e
X Th

Considering now the deformation gradient tensor and the fundamental re-
er
tin

lation dx = F · dX defined in (2.2) as well as the polar decomposition given


in (2.49), the following is obtained9 .
on

.O

stretching
  
C

rotation
  
©

dx = F · dX = (V · Q) · dX = V · (Q · dX) (2.50)
not
F (•) ≡ stretching ◦ rotation (•)

7 A second-order tensor Q is orthogonal if QT · Q = Q · QT = 1 is verified.


8 To obtain the square root of a tensor, first the tensor must be diagonalized, then the square
root of the elements in the diagonal of the diagonalized component matrix are obtained and,
finally, the diagonalization is undone.
9 The notation (◦) is used here to indicate the composition of two operations ξ and ϕ:
z = ϕ ◦ ξ (x).

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Polar Decomposition 63

rotation
  
stretching
  
dx = F · dX = (Q · U) · dX = Q · (U · dX) (2.51)
not
F (•) ≡ rotation ◦ stretching (•)

Remark 2.11. An orthogonal tensor Q (such that |Q| = 1) is named

rs
rotation tensor and the mapping y = Q · x is denominated rotation.

ee
A rotation has the following properties:

s gin
• When applied on any vector x, the result is another vector
y = Q · x with the same modulus,

t d le En

y
2 = y·y ≡ [y]T ·[y] = [Q · x]T ·[Q · x] ≡ x·QT · Q ·x = x·x =
x
2 .
not not

  

r
ba
ge ro or
eS m
1

ci
f

ra
• The result of multiplying (mapping) the orthogonal tensor Q to
C d P cs
b
a
two vectors x(1) and x(2) with the same origin and that form an
i
an an n

angle α between them, maintains the same angle between the


y ha

images y(1) = Q · x(1) and y(2) = Q · x(2) ,


le
liv or ec

y(1) · y(2) x(1) · QT · Q · x(2) x(1) · x(2)


( ( ( ( = ( ( ( ( = ( ( ( ( = cos α .
M

.A

(y(1) ( (y(2) ( (y(1) ( (y(2) ( (x(1) ( (x(2) (


m

d
uu

In consequence, the mapping (rotation) y = Q · x maintains the an-


e
X Th

er

gles and distances.


tin
on

.O
C

Remark 2.12. Equations (2.50) establish that the relative motion in


the neighborhood of the particle during the deformation process
(characterized by tensor F ) can be understood as the composition
of a rotation (characterized by the rotation tensor Q, which main-
tains angles and distances) and a stretching or deformation in itself
(which modifies angles and distances) characterized by the tensor V
(see Figure 2.10).

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
64 C HAPTER 2. S TRAIN

rs
ee
s gin
t d le En
Figure 2.10: Polar decomposition.

r
ba
ge ro or
eS m
ci
f

ra
C d P cs
b
a
i
an an n

Remark 2.13. Alternatively, equations (2.51) allow characterizing


y ha

the relative motion in the neighborhood of a particle during the de-


le
formation process as the superposition of a stretching or deformation
liv or ec

in itself (characterized by tensor U ) and a rotation (characterized by


M

.A

the rotation tensor Q ).


m

A rigid body motion is a particular case of deformation characterized


d

by U = V = 1 and Q = F.
uu
e
X Th

er
tin
on

.O

2.9 Volume Variation


C

Consider a particle P of the continuous medium in the reference configuration


(t = 0) which has a differential volume dV0 associated with it (see Figure 2.11).
This differential volume is characterized by the positions of another three par-
ticles Q, R and S belonging to the differential neighborhood of P, which are
aligned with this particle in three arbitrary directions. The volume differential
dVt , associated with the same particle in the present configuration (at time t),
   
will also be characterized by the spatial points P , Q , R and S corresponding
to Figure 2.11 (the positions of which define a parallelepiped that is no longer
oriented along the coordinate axes).
The relative position vectors between the particles in the material configura-
tion are dX(1) , dX(2) and dX(3) , and their counterparts in the spatial configura-

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Volume Variation 65

tion are dx(1) = F · dX(1) , dx(2) = F · dX(2) and dx(3) = F · dX(3) . Obviously, the
relations 
dx(i) = F · dX(i)
(i) (i) (2.52)
dx j = Fjk dXk i, j, k ∈ {1, 2, 3}
are satisfied. Then, the volumes10 associated with a particle in both configura-
tions can be written as
⎡ (1) (1) (1)

dX1 dX2 dX3
  ⎢ (2) ⎥
dV0 = dX(1) × dX(2) · dX(3) = det ⎢ (2) (2) ⎥
⎣ dX1 dX2 dX3 ⎦ = |M| ,

rs
ee
(3) (3) (3)
dX1 dX2 dX3
  

s gin
[M]

t d le En
⎡ (1) (1) (1)

r
dx1 dx2 dx3
 

ba
ge ro or
⎢ (2) ⎥

eS m
dVt = dx(1) × dx(2) · dx(3) = det ⎢ (2) (2) ⎥
⎣ dx1 dx2 dx3 ⎦ = |m| ,

ci
(2.53)
f

ra
C d P cs
(3) (3) (3)

b
a
dx1 dx2 dx3
  
i
an an n

[m]
y ha

le
liv or ec

(i) (i)
where Mi j = dX j and mi j = dx j . Considering these expressions,
M

.A

(i) (i)
mi j = dx j = Fjk dXk = Fjk dMik = dMik FkTj =⇒ m = M · FT (2.54)
m

d
uu
e

is deduced and, consequently11 ,


X Th

er
tin

" " " " ⎫


dVt = |m| = "M · FT " = |M| "FT " = |F| |M| = |F| dV0 ⎪ ⎪

on

.O


dV0 =⇒ dVt = |F|t dV0

C


dVt = dV (x (X,t) ,t) = |F (X,t)| dV0 (X, 0) = |F|t dV0 ⎭
©

(2.55)

10 The volume of a parallelepiped is calculated as the scalar triple product (a × b) · c of the


concurrent edge-vectors a, b and c, which meet at any of the parallelepiped’s vertices. Note
that the scalar triple product is the determinant of the matrix constituted by the components
of the above mentioned vectors arranged "in rows.
"
11 The expressions |A · B| = |A| · |B| and "AT " = |A| are used here.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
66 C HAPTER 2. S TRAIN

rs
ee
s gin
Figure 2.11: Variation of a volume differential element.

t d le En

r
2.10 Area Variation

ba
ge ro or
eS m
ci
f
Consider an area differential dA associated with a particle P in the reference

ra
C d P cs
configuration and its variation along time. To define this area differential, con-
b
a
i
sider two particles Q and R in the differential neighborhood of P, whose relative
an an n

positions with respect to this particle are dX(1) and dX(2) , respectively (see Fig-
y ha

ure 2.12). Consider also an arbitrary auxiliary particle S whose relative position
le
liv or ec

vector is dX(3) . An area differential vector dA = dA N associated with the scalar


M

.A

differential area, dA, is defined. The module of vector dA is dA and its direction
is the same as that of the unit normal vector in the material configuration N.
m

In the present configuration, at time t, the particle will occupy a point in


d
uu


e

space P and will have an area differential da associated with it which, in turn,
X Th

defines an area differential vector da = da n, where n is the corresponding unit


er
tin

normal vector in the spatial configuration. Consider also the positions of the
  
other particles Q , R and S and their relative position vectors dx(1) , dx(2) and
on

.O

dx(3) .
C

The volumes dV0 and dVt of the corresponding parallelepipeds can be calcu-
©

lated as
dV0 = dH dA = dX(3) · N dA = dX(3) · N dA = dA · dX(3)
   
dH dA (2.56)
(3)
dVt = dh da = dx · n da = dx · n da = da · dx(3)
(3)
   
dh da

and, taking into account that dx(3) = F · dX(3) , as well as the expression for
change in volume (2.55), results in
da · F · dX(3) = da · dx(3) = dVt = |F| dV0 = |F| dA · dX(3) ∀dX(3) . (2.57)

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Infinitesimal Strain 67

rs
ee
Figure 2.12: Variation of an area differential.

s gin
t d le En
Comparing the first and last terms12 in (2.57) and considering that the relative

r
position of particle S can take any value (as can, therefore, vector dX(3) ), finally

ba
ge ro or
eS m
ci
yields
f

ra
da · F = |F| dA =⇒ da = |F| dA · F−1 .
C d P cs
(2.58)
b
a
i
an an n

To obtain the relation between the two area differential scalars, dA and da,
y ha

expressions dA = N dA and da = n da are replaced into (2.58) and the modules


le
are taken, resulting in
liv or ec

( (
M

.A

da n = |F| N · dF−1 dA =⇒ da = |F| (N · dF−1 ( dA . (2.59)


m

2.11 Infinitesimal Strain


uu
e
X Th

er
tin

Infinitesimal strain theory (also denominated small deformation theory) is based


on two simplifying hypotheses of the general theory (or finite strain theory) seen
on

.O

in the previous sections (see Figure 2.13).


C

Definition 2.5. The simplifying hypotheses are:


1) Displacements are very small compared to the typical dimensions
in the continuous medium:
u
<<
X
.
2) Displacement gradients are very small (infinitesimal).

12 Here, the following tensor algebra theorem is taken into account: given two vectors a and
b, if the relation a · x = b · x is satisfied for all values of x, then a = b.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
68 C HAPTER 2. S TRAIN

rs
ee
s gin
Figure 2.13: Infinitesimal strain in the continuous medium.

t d le En

r
ba
ge ro or
eS m
In accordance with the first hypothesis, the reference configuration Ω0 and

ci
f
the present configuration Ωt are very close together and are considered to be

ra
C d P cs
b
indistinguishable from one another. Consequently, the material and spatial co-

a
i
ordinates coincide and discriminating between material and spatial descriptions
an an n
y ha

no longer makes sense.


⎧ ⎧
le
liv or ec

⎨x = X+u ∼ = X ⎨ U (X,t) not


= u (X,t) ≡ u (x,t)
=⇒
M

.A

⎩ xi = Xi + ui ∼
= Xi ⎩ Ui (X,t) not
= ui (X,t) ≡ ui (x,t) i ∈ {1, 2, 3}
m

(2.60)
d
uu

The second hypothesis can be written in mathematical form as


e
X Th

" "
er
tin

" ∂ ui "
" " ∀ i, j ∈ {1, 2, 3} .
"∂xj " 1 (2.61)
on

.O
C

2.11.1 Strain Tensors. Infinitesimal Strain Tensor


The material and spatial displacement gradient tensors coincide. Indeed, in view
of (2.60),

⎨x = X ∂ ui ∂Ui
j j
=⇒ ji j = = = Ji j =⇒ j = J (2.62)
⎩ ui (x,t) = Ui (X,t) ∂xj ∂ Xj

and the material strain tensor results in

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Infinitesimal Strain 69



⎪ 1  
T + JT · J ∼ 1 J + JT ,


⎪ E = J + J =

⎨ 2 2
 
1 ∂ ui ∂ u j ∂ uk ∂ uk ∼ 1 ∂ ui ∂ u j (2.63)

⎪ Ei j = + + = + ,

⎪ 2 ∂ x j ∂ xi ∂ xi ∂ x j 2 ∂ x j ∂ xi

⎩   
1
where the infinitesimal character of the second-order term (∂ uk ∂ uk /∂ x j ∂ xi ) has
been taken into account. Operating in a similar manner with the spatial strain
tensor,

rs

⎪ 1    
T − jT · j ∼ 1 j + jT = 1 J + JT ,


ee

⎪ e = j + j =
⎨ 2

2 2


s gin
1 ∂ ui ∂ u j ∂ uk ∂ uk ∼ 1 ∂ ui ∂ u j (2.64)

⎪ e ij = + − = + .

⎪ ∂ ∂ ∂ ∂ ∂ ∂

t d le En
⎪ 2 x j x i x i x j 2 x j xi
⎩   
1

r
ba
ge ro or
eS m
ci
Equations (2.63) and (2.64) allow defining the infinitesimal strain tensor (or
f

ra
small strain tensor) ε as13
C d P cs
b
a

i
an an n


⎪ 1  not
y ha


⎨ε = J + JT = ∇s u
le
Infinitesimal 2

liv or ec

(2.65)
strain tensor ⎪
⎪ 1 ∂ ui ∂ u j
⎪ ε
⎩ ij = + i, j ∈ {1, 2, 3}
M

.A

2 ∂ x j ∂ xi
m

d
uu
e
X Th

er
tin

Remark 2.14. Under the infinitesimal strain hypothesis, the material


and spatial strain tensors coincide and collapse into the infinitesimal
on

.O

strain tensor.
E (x,t) = e (x,t) = ε (x,t)
C

Remark 2.15. The infinitesimal strain tensor is symmetric, as ob-


served in its definition in (2.65).
1 T 1  T 
εT = J + JT = J +J = ε
2 2

13 The symmetric gradient operator ∇s is defined as ∇s (•) = ((•) ⊗ ∇ + ∇ ⊗ (•)) /2.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
70 C HAPTER 2. S TRAIN

Remark 2.16. The components of the infinitesimal strain tensor ε are


infinitesimal (εi j 1). Proof is obvious from (2.65) and the condi-
tion that the components of J = j are infinitesimal (see (2.61)).

Example 2.4 – Determine under which conditions the motion in Example 2.1
constitutes an infinitesimal strain case and obtain the infinitesimal strain ten-

rs
sor for this case. Compare it with the result obtained from the spatial and

ee
material strain tensors in Example 2.2 taking into account the infinitesimal
strain hypotheses.

s gin
t d le En
Solution

r
The equation of motion is given by

ba
ge ro or
eS m

ci
⎨ x1 = X1 − AX3
⎪ f

ra
C d P cs
b
a
x2 = X2 − AX3 ,

i

an an n

x3 = −AX1 + AX2 + X3
y ha

le
liv or ec

from which the displacement field is obtained


M

⎡ ⎤
.A

U1 = −AX3
not ⎢ ⎥
m

U (X,t) = x − X ≡ ⎣ U2 = −AX3 ⎦.
d
uu
e

U3 = −AX1 + AX2
X Th

er
tin

It is obvious that, for the displacements to be infinitesimal, A must be in-


on

.O

finitesimal (A 1). Now, to obtain the infinitesimal strain tensor, first the
displacement gradient tensor J (X,t) = j (x,t) must be computed,
C

⎡ ⎤ ⎡ ⎤
−AX3   0 0 −A
not ⎢ ⎥ ∂ ∂ ∂ ⎢ ⎥
J = U⊗∇ ≡ ⎢ ⎣ −AX3 ⎥
⎦ ∂ X1 , ∂ X2 , ∂ X3 = ⎣ 0 0 −A ⎦ .
−AX1 + AX2 −A A 0

Then, the infinitesimal strain tensor, in accordance to (2.65), is


⎡ ⎤
0 0 −A
not ⎢ ⎥
ε = ∇s U ≡ ⎣ 0 0 0 ⎦ .
−A 0 0

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Infinitesimal Strain 71

The material and spatial strain tensors obtained in Example 2.2 are, respec-
tively,
⎡ 2 ⎤
A −A2 −2A
not 1 ⎢ ⎥
E (X,t) ≡ ⎣ −A2 A2 0 ⎦ and
2
−2A 0 2A2
⎡ ⎤
−3A2 − 2A4 A2 + 2A4 −2A − 2A3
not 1 ⎢ ⎥
e (X,t) ≡ ⎣ A2 + 2A4 A2 − 2A4 2A3 ⎦.
2
−2A − 2A 3 3 −2A 2

rs
2A

ee
Neglecting
 4 the second-order
 and higher-order infinitesimal terms
A A3 A2 A results in

s gin
⎡ ⎤ ⎡ ⎤

t d le En
0 0 −A 0 0 −A
not ⎢ ⎥ not ⎢ ⎥
0 −A ⎦ =⇒ E = e = ε ,

r
E≡⎣ 0 0 −A ⎦ and e≡⎣ 0

ba
ge ro or
eS m
−A A −A A

ci
0 0
f

ra
C d P cs
b
a
which is in accordance with Remark 2.14.
i
an an n
y ha

2.11.2 Stretch. Unit Elongation


le
liv or ec

Considering
 the general expression
√  (2.30) of the unit elongation in the direction
M

.A

T∼ = t λt = 1 + 2t · E · t and applying a Taylor series expansion14 around 0


(taking into account that E = ε is infinitesimal and, therefore, so is x = t · ε · t ),
m

yields
uu
e


X Th

er

1 + 2t · ε · t ∼
tin

λt = = 1+t·ε ·t
   (2.66)
on

x
.O

εt = λt − 1 = t · ε · t
C

2.11.3 Physical Interpretation of the Infinitesimal Strains


Consider the infinitesimal strain tensor ε and its components in the coordinate
system x1 ≡ x, x2 ≡ y, x3 ≡ z, shown in Figure 2.14,
⎡ ⎤ ⎡ ⎤
εxx εxy εxz ε11 ε12 ε13
not ⎢ ⎥ ⎢ ⎥
ε ≡ ⎣ εxy εyy εyz ⎦ = ⎣ ε12 ε22 ε23 ⎦ . (2.67)
εxz εyz εzz ε13 ε23 ε33
√ √  
14 The Taylor series expansion of 1 + x around x = 0 is 1 + x = 1 + x/2 + O x2 .

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
72 C HAPTER 2. S TRAIN

Figure 2.14: Physical interpretation of the infinitesimal strains.

rs
ee
Consider a differential segment PQ oriented in the reference configuration

s gin
parallel to the coordinate axis x1 ≡ x. The stretch λx and the unit elongation εx
in this direction are, according to (2.66) with t = [1, 0, 0]T ,

t d le En
λx = 1 + t · ε · t = 1 + εxx =⇒ εx = λx − 1 = εxx .

r
(2.68)

ba
ge ro or
eS m
ci
This allows assigning to the component εxx ≡ ε11 the physical meaning of unit
f

ra
elongation εx in the direction of the coordinate axis x1 ≡ x. A similar interpre-
C d P cs
b
a
tation is deduced for the other components in the main diagonal of the tensor
i
an an n

ε (εxx , εyy , εzz ),


y ha

εxx = εx ; εyy = εy ; εzz = εz . (2.69)


le
liv or ec

Given now the components outside the main diagonal of ε , consider the dif-
M

.A

ferential segments PQ and PR oriented in the reference configuration parallel to


m

the coordinate directions x and y, respectively. Then, these two segments form
d

an angle Θxy = π/2 in this configuration. Applying (2.43), the increment in the
uu
e

corresponding angle results in15


X Th

er
tin

π εxy ∼
Δ θxy = θxy − = −2 arcsin $ $ = −2 arcsin εxy = −2εxy ,
on

.O

2 1 + 2εxx 1 + 2εyy   
       εxy
C

1 1
©

(2.70)
where the infinitesimal character of εxx , εyy and εxy has been taken into account.
Consequently, εxy can be interpreted from (2.70) as minus the semi-increment,
produced by the strain, of the angle between the two differential segments ini-
tially oriented parallel to the coordinate directions x and y. A similar interpre-
tation is deduced for the other components εxz and εyz ,

1 1 1
εxy = − Δ θxy ; εxz = − Δ θxz ; εyz = − Δ θyz . (2.71)
2 2 2
 
15 The Taylor series expansion of arcsin x around x = 0 is arcsin x = x + O x2 .

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Infinitesimal Strain 73

2.11.4 Engineering Strains. Vector of Engineering Strains


There is a strong tradition in engineering to use a particular denomination for
the components of the infinitesimal strain tensor. This convention is named en-
gineering notation, as opposed to the scientific notation generally used in con-
tinuum mechanics. Both notations are synthesized as follows.

engineering notation
 ⎡  
scientific notation
  ⎤
⎡ ⎤ ⎡ ⎤
ε11 ε12 ε13 εxx εxy εxz εx 2 γxy 2 γxz
1 1

⎢ ⎥ ⎢ ⎥ ⎢ ⎥
ε22 ε23 ⎦ ≡ ⎣ εxy εyy εyz ⎦ ≡ ⎢ ⎥
not
ε ≡ ⎣ ε12

rs
⎣2
1
γxy ε y
1
2 γyz ⎦ (2.72)

ee
ε13 ε23 ε33 εxz εyz εzz
2 γxz 2 γyz εz
1 1

s gin
t d le En

r
Remark 2.17. The components in the main diagonal of the strain ten-

ba
ge ro or
eS m
sor (named longitudinal strains) are denoted by ε(•) and coincide

ci
f

ra
with the unit elongations in the directions
 of the  coordinate axes.
C d P cs
Positive values of longitudinal strains ε(•) > 0 correspond to an
b
a
i
increase in length of the corresponding differential segments in the
an an n
y ha

reference configuration.
le
liv or ec
M

.A
m

Remark 2.18. The components outside the main diagonal of the


uu

strain tensor are characterized by the values γ(•, •) (named angu-


e
X Th

er

lar strains) and can be interpreted as the decrements of the corre-


tin

sponding angles oriented in the Cartesian directions


 of the reference
configuration. Positive values of angular strains γ(•, •) > 0 indicate
on

.O

that the corresponding angles close with the deformation process.


C

In engineering, it is also frequent to exploit the symmetry of the infinitesimal


strain tensor (see Remark 2.15) to work only with the six components of the
tensor that are different, grouping them in the vector of engineering strains,
which is defined as follows.

de f
 T
ε ∈ R6 ε = εx , εy , εx , γxy , γxz , γyz
      (2.73)
longitudinal angular
strains strains

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
74 C HAPTER 2. S TRAIN

2.11.5 Variation of the Angle between Two Differential Segments in


Infinitesimal Strain
Consider any two differential segments, PQ and PR, in the reference configu-
ration and the angle Θ they define (see Figure 2.15). The angle formed by the
corresponding deformed segments in the present configuration is θ = Θ + Δ θ .
Applying (2.42) to this case results in

T(1) · (1 + 2εε ) · T(2)


cos θ = cos (Θ + Δ θ ) = $ $ , (2.74)
1 + 2T(1) · ε · T(1) 1 + 2T(2) · ε · T(2)
     

rs
1 1

ee
where T(1) and T(2) are the unit vectors

s gin
( ( (in the( directions of PQ and PR and,
(1) (2) ( (1) ( ( (2) (
therefore, the relation T · T = (T ( (T ( cosΘ = cosΘ is fulfilled. Con-

t d le En
sidering the infinitesimal character of the components of ε and Δ θ , the follow-
ing holds true16 .

r
ba
ge ro or
eS m
ci
f
cos θ = cos (Θ + Δ θ ) = cosΘ · cos Δ θ − sinΘ · sin Δ θ =

ra
     
C d P cs
b
a
≈1 ≈ Δθ
i
an an n
y ha

= cosΘ
  
le
liv or ec

T(1) · T(2) +2T(1) · ε · T(2) (2.75)


= cosΘ − sinΘ · Δ θ = $ $ =
M

.A

1 + T(1) · ε · T(1) 1 + T(2) · ε · T(2)


     
m

≈1 ≈1
d
uu

= cosΘ + 2T(1) · ε · T(2)


e
X Th

er
tin

Therefore, sinΘ · Δ θ = −2T(1) · ε · T(2) and


on

.O

2T(1) · ε · T(2) 2t(1) · ε · t(2)


C

Δθ = − =− , (2.76)
©

sinΘ sin θ

where the infinitesimal character of the strain has been taken into account and,
thus, it follows that T(1) ≈ t(1) , T(2) ≈ t(2) and Θ ≈ θ .

2.11.6 Polar Decomposition


The polar decomposition of the deformation gradient tensor F is given by (2.49)
for the general case of finite strain. In the case of infinitesimal strain, recall-
 
16The following Taylor series expansions around x = 0 are considered: sin x = x + O x2
 2
and cos x = 1 + O x

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Infinitesimal Strain 75

rs
ee
Figure 2.15: Variation of the angle between two differential segments in infinitesimal
strain.

s gin
t d le En
ing (2.12) and the infinitesimal character of the components of the tensor J

r
(see (2.61)), tensor U in (2.49) can be written as17

ba
ge ro or
eS m
$ $ ⎫

ci
U = FT · F = (1 + JT ) · (1 + J) = f ⎪

ra

C d P cs


b
a
$ $  
i
1
= 1 + J + J + J · J ≈ 1 + J + J = 1 + J + J ⎪ =⇒ U = 1 + ε .
an an n

T T T T
   ⎪
y ha

   ⎪
2

J ε
le
liv or ec

(2.77)
M

.A

In a similar manner, due to the infinitesimal character of the components of the


tensor ε (see Remark 2.16), the inverse of tensor U results in18
m

1 
uu

U−1 = (1 + ε )−1 = 1 − ε = 1 −
e

J + JT . (2.78)
X Th

2
er
tin

Therefore, the rotation tensor Q in (2.49) can be written as


on

.O

 ⎫
1  ⎪
C

−1
Q = F · U = (1 + J) · 1 − J + J T = ⎪



©

2 ⎬
1  1   1   =⇒ Q = 1 + Ω .
= 1 + J − J + JT − J · J + JT = 1 + J − JT ⎪ ⎪

2 2   2   ⎪ ⎪

J Ω
(2.79)

√ √  
17 The Taylor series expansions of tensor 1 + x around x = 0 is 1 + x = 1 + x/2 + O x2 .
18 The Taylor series expansions of tensor (1 + x)−1 around x = 0 is (1 + x)−1 = 1 − x +
 
O x2 .

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
76 C HAPTER 2. S TRAIN

Equation (2.79) defines the infinitesimal rotation tensor Ω 19 as follows.




⎪ de f 1   1

⎨Ω =
de f
J − JT = (u ⊗ ∇ − ∇ ⊗ u) = ∇a u
Infinitesimal 2 2
 (2.80)
rotation tensor ⎪
⎪ 1 ∂ ui ∂ u j

⎩ Ωi j = − 1 i, j ∈ {1, 2, 3}
2 ∂ x j ∂ xi

Remark 2.19. The tensor Ω is antisymmetric. Indeed,

rs

ee
⎨ Ω T = 1 J − JT T = 1 JT − J = −Ω Ω

s gin
2 2 .
⎩ Ω = −Ω i, j ∈ {1, 2, 3}
ji ij

t d le En
Consequently, the terms in the main diagonal of Ω are zero, and its

r
ba
ge ro or
eS m
matrix of components has the structure

ci
⎡ f ⎤

ra
C d P cs
0 Ω12 −Ω31
b
a
Ω] = ⎣ −Ω12 Ω23 ⎦ .
i
[Ω 0
an an n
y ha

Ω31 −Ω23 0
le
liv or ec
M

.A

In a small rotation context, tensor Ω characterizes the rotation (Q = 1 + Ω )


m

and, thus, the denomination of infinitesimal rotation tensor. Since it is an anti-


uu

symmetric tensor, it is defined solely by three different components (Ω23 , Ω31 ,


e
X Th

Ω12 ), which form the infinitesimal rotation vector θ 20 ,


er
tin

⎡ ⎤
on

.O

Infinitesimal ∂ u 3 ∂ u2
rotation vector: ⎡ ⎤ ⎡ ⎤ ⎢ − ⎥
C

θ1 −Ω23 ⎢ ∂ x2 ∂ x3 ⎥
⎢ ⎥ 1
©

not ⎢ ⎥ ⎢ ⎥ 1⎢ ⎥ de f
θ ≡ ⎣ θ2 ⎦ = ⎣ −Ω31 ⎦ = ⎢ ∂ u1 − ∂ u3 ⎥ = ∇ × u . (2.81)
2 ⎢ ∂ x3 ∂ x1 ⎥ 2
θ3 −Ω12 ⎢ ⎥
⎣ ∂u ∂ u1 ⎦
2

∂ x1 ∂ x2

19 The antisymmetric gradient operator ∇a is defined as ∇a (•) = [(•) ⊗ ∇ − ∇ ⊗ (•)] /2.


20 The operator rotational of (•) is denoted as ∇ × (•).

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Infinitesimal Strain 77

Expressions (2.12), (2.65) and (2.79) allow writing


1  1 
F = 1+J+ J + JT + J − JT =⇒ F = 1 + ε + Ω . (2.82)
2 2
     
ε Ω

Remark 2.20. The results of performing a dot product of the in-


finitesimal rotation tensor Ω and performing a cross product of the
infinitesimal rotation vector θ with any vector r ≡ [r1 , r2 , r3 ]T (see

rs
Figure 2.16) coincide. Indeed,

ee
⎡ ⎤⎡ ⎤ ⎡ ⎤

s gin
0 Ω12 −Ω31 r1 Ω12 r2 − Ω31 r3
not ⎢ ⎥⎢ ⎥ ⎢ ⎥
Ω · r ≡ ⎣ −Ω12 0 Ω23 ⎦ ⎣ r2 ⎦ = ⎣ −Ω12 r1 + Ω23 r3 ⎦ ,

t d le En
Ω31 −Ω23 0 r3 Ω31 r1 − Ω23 r2

r
ba
ge ro or
⎡ ⎤ ⎡ ⎤

eS m
ci
ê1 ê2 ê3 ê1 ê2 ê3
f

ra
not ⎢ ⎥ ⎢ ⎥
C d P cs
θ × r ≡ ⎣ θ1 θ2 θ3 ⎦ = ⎣ −Ω23 −Ω31 −Ω12 ⎦ =
b
a
i
an an n

r1 r2 r3 r1 r2 r3
⎡ ⎤
y ha

Ω12 r2 − Ω31 r3
le
⎢ ⎥
liv or ec

= ⎣ −Ω12 r1 + Ω23 r3 ⎦ .
M

.A

Ω31 r1 − Ω23 r2
m

Consequently, vector Ω · r = θ × r has the following characteristics:


uu
e
X Th

• It is orthogonal to vector r (because it is the result of a vector


er
tin

product in which r is involved).


on

.O

• Its module is infinitesimal (because θ is infinitesimal).


C

• Vector r + Ω · r = r + θ × r can be considered, except for higher-


©

order infinitesimals, as the result of applying a rotation θ on


vector r.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
78 C HAPTER 2. S TRAIN

rs
ee
Figure 2.16: Product of the infinitesimal rotation vector and tensor on a vector r.

s gin
t d le En
Consider now a differential segment dX in the neighborhood of a particle P
in the reference configuration (see Figure 2.17). In accordance with (2.82), the

r
ba
ge ro or
stretching transforms this vector into vector dx as follows.

eS m
ci
f

ra
stretching rotation
C d P cs
     
b
a
dx = F · dX = (1 + ε + Ω) · dX = ε · dX + (1 + Ω) · dX
i
an an n

(2.83)
y ha

le
F (•) ≡ stretching (•) + rotation (•)
liv or ec
M

.A
m

d
uu

Remark 2.21. Under infinitesimal strain hypotheses, the expression


e
X Th

in (2.83) characterizes the relative motion of a particle, in the differ-


er
tin

ential neighborhood of this particle, as the following sum:


on

.O

a) A stretching or deformation in itself, characterized by the in-


finitesimal strain tensor ε .
C

b) A rotation characterized by the infinitesimal rotation tensor Ω


which, in the infinitesimal strain context, maintains angles and
distances.
The superposition (stretching ◦ rotation) of the general finite strain
case (see Remark 2.12) degenerates, for the infinitesimal strain case,
into a simple addition (stretching + rotation).

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Volumetric Strain 79

Figure 2.17: Stretching and rotation in infinitesimal strain.

rs
ee
s gin
t d le En
2.12 Volumetric Strain

r
ba
ge ro or
eS m
ci
f

ra
C d P cs
Definition 2.6. The volumetric strain is the increment produced by
b
a
i
the deformation of the volume associated with a particle, per unit of
an an n

volume in the reference configuration.


y ha

le
liv or ec
M

.A

This definition can be mathematically expressed as (see Figure 2.18)


m

de f dV (X,t) − dV (X, 0) not dVt − dV0


uu

Volumetric strain: e (X,t) = = . (2.84)


e

dV (X, 0) dV0
X Th

er
tin
on

.O
C

Figure 2.18: Volumetric strain.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
80 C HAPTER 2. S TRAIN

Equation (2.55) allows expressing, in turn, the volumetric strain as follows:


• Finite strain

dVt − dV0 |F|t dV0 − dV0


e= = =⇒ e = |F| − 1 (2.85)
dV0 dV0

• Infinitesimal strain

Considering (2.49) and recalling that Q is an orthogonal tensor (|Q| = 1), yields

rs
⎡ ⎤

ee
1 + εxx εxy εxz
⎢ ⎥

s gin
|F| = |Q · U| = |Q| |U| = |U| = |1 + ε | = det ⎣ εxy 1 + εyy εyz ⎦ ,
εxz εyz 1 + εzz

t d le En
(2.86)

r
ba
where (2.77) has been considered. Taking into account that the components of ε

ge ro or
eS m
ci
are infinitesimal, and neglecting in the expression of its determinant the second-
f

ra
order and higher-order infinitesimal terms, results in
C d P cs
b
a
⎡ ⎤
i
an an n

1 + εxx εxy εxz


 
y ha

⎢ ⎥
|F| = det ⎣ εxy 1 + εyy εyz ⎦ = 1 + εxx + εyy + εzz +O ε 2 ≈ 1 + Tr (εε ) .
le
  
liv or ec

εxz εyz 1 + εzz Tr (εε )


M

.A

(2.87)
m

Then, introducing (2.87) into (2.85) yields, for the infinitesimal strain case
d


uu
e

dVt = (1 + Tr (εε )) dV0 ⎪⎬


X Th

er
tin

dVt − dV0 =⇒ e = Tr (εε ) . (2.88)


e= = |F| − 1 ⎪

on

.O

dV0
C

2.13 Strain Rate


In the previous sections of this chapter, the concept of strain has been studied,
defined as the variation of the relative position (angles and distances) of the
particles in the neighborhood of a given particle. In the following sections, the
rate at which this relative position changes will be considered by introducing
the concept of strain rate as a measure of the variation in the relative position
between particles per unit of time.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Strain Rate 81

2.13.1 Velocity Gradient Tensor


Consider the configuration corresponding to a time t, two particles of the con-
tinuous medium P and Q that occupy the spatial points P and Q at said instant
of time (see Figure 2.19), their velocities vP = v (x,t) and vQ = v (x + dx,t), and
their relative velocity,
dv (x,t) = vQ − vP = v (x + dx,t) − v (x,t) . (2.89)

Then, ⎧

⎪ ∂v

rs
⎨ dv = · dx = l · dx
∂x

ee
, (2.90)


⎪ ∂ vi

s gin

⎩ dvi = dx j = li j dx j i, j ∈ {1, 2, 3}
∂xj

t d le En
where the spatial velocity gradient tensor l (x,t) has been introduced.

r
ba

ge ro or
eS m
⎪ de f ∂ v (x,t)

ci
⎪ l (x,t) =

⎪ f

ra

⎨ ∂x
C d P cs
b
a
Spatial velocity
l = v⊗∇
i
(2.91)
gradient tensor ⎪
an an n


⎪ ∂ vi
y ha


⎪ i, j ∈ {1, 2, 3}
⎩ li j =
∂xj
le
liv or ec
M

.A
m

d
uu
e
X Th

er
tin
on

.O
C

Figure 2.19: Velocities of two particles in the continuous medium.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
82 C HAPTER 2. S TRAIN

2.13.2 Strain Rate and Spin Tensors


The velocity gradient tensor can be split into a symmetric and an antisymmetric
part21 ,
l = d+w , (2.92)
where d is a symmetric tensor denominated strain rate tensor,


⎪ de f 1  1

⎪ d = sym (ll ) = l + l T not
= (v ⊗ ∇ + ∇ ⊗ v) = ∇s v

⎪ 2 2

⎪ 

⎨ d = 1 ∂ vi + ∂ v j
Strain ⎪

rs
ij i, j ∈ {1, 2, 3}
2 ∂ x j ∂ xi

ee
rate (2.93)

tensor ⎪


⎡ ⎤

s gin

⎪ d 11 d 12 d 13



⎪ [d] = ⎣ d12 d22 d23 ⎦

t d le En
d13 d23 d33

r
ba
ge ro or
eS m
ci
and w is an antisymmetric tensor denominated rotation rate tensor or spin ten-
f

ra
sor, whose expression is
C d P cs
b
a
i

an an n


⎪ 1  1
y ha

de f

⎪ w = skew (l
l ) = l − l T not
= (v ⊗ ∇ − ∇ ⊗ v) = ∇a v


le
⎪ 2 2

liv or ec

Rotation⎪

⎨ w = 1 ∂ vi − ∂ v j

i, j ∈ {1, 2, 3}
M

.A

rate ij
2 ∂ x j ∂ xi (2.94)
(spin) ⎪
⎪ ⎡ ⎤
tensor ⎪
m


⎪ 0 w12 −w31
d



uu


⎪ [w] = ⎣ −w12 0 w23 ⎦
e



X Th

er

w31 −w23 0
tin
on

.O
C

2.13.3 Physical Interpretation of the Strain Rate Tensor


©

Consider a differential segment defined by the particles P and Q of Figure 2.20


and the variation of their squared length along time,
d 2 d d d
ds = (dx · dx) = (dx) · dx + dx · (dx) =
dt dt dt  dt
dx dx
=d · dx + dx · d = dv · dx + dx · dv , (2.95)
dt dt
21 Every second-order tensor a can be decomposed into the sum of its symmetric part
(sym (a)) and its antisymmetric or skew-symmetric part (skew (a)) in the form:
a = sym (a) + skew (a) with sym (a) = (a + aT )/2 and skew (a) = (a − aT )/2.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Strain Rate 83

and using relations (2.90) and (2.93), the expression


d 2    
ds = dx · l T · dx + dx · (ll · dx) = dx · l T + l · dx = 2 dx · d · dx (2.96)
dt
is obtained. Differentiating now (2.20) with respect to time and taking into ac-
count (2.96) yields
d d  2 
2 dx · d · dx = ds2 (t) = ds (t) − dS2 =
dt dt (2.97)
d dE .
= (2 dX · E (X,t) · dX) = 2 dX · · dX = 2 dX · E · dX .

rs
dt dt

ee
Replacing (2.2) into (2.97) results in22

s gin
.  
dX · E · dX = dx · d · dx ≡ [dx]T [d] [dx] = [dX]T FT · d · F [dX]
not

t d le En
 . .
=⇒ dX · FT · d · F − E · dX = 0 ∀ dX =⇒ FT · d · F − E = 0

r
ba
ge ro or
eS m
.

ci
E = FT · d · F .
f (2.98)

ra
C d P cs
b
a
i
an an n
y ha

Remark 2.22. Equation (2.98) shows the existing relationship be-


tween the strain rate tensor d (x,t) and the material derivative of
le
liv or ec

.
the material strain tensor E (X,t), providing a physical interpreta-
M

.A

tion (and justifying the denomination) of tensor d (x,t). However,


.
the same equation reveals that tensors d (x,t) and E (X,t) are not
m

exactly the same. Both tensors will coincide in the following cases:
uu

"
e

• In the reference configuration: t = t0 ⇒ F"


X Th

= 1.
er
tin

t=t0
∂x
on

• In infinitesimal strain theory: x ≈ X ⇒ F = ≈ 1.


.O

∂X
C

22 Here, the following tensor algebra theorem is used: given a second-order tensor A, if
x · A · x = 0 is verified for all vectors x = 0, then A ≡ 0.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
84 C HAPTER 2. S TRAIN

Figure 2.20: Differential segment between particles of the continuous medium along

rs
time.

ee
s gin
2.13.4 Physical Interpretation of the Rotation Rate Tensor

t d le En
Taking into account the antisymmetric character of w (which implies it can be

r
defined using only three different components), the vector

ba
ge ro or
eS m
⎡  ⎤

ci
∂ v 2 ∂ v3 f

ra
⎢−
C d P cs
− ⎥ ⎡ ⎤
b
a
⎢ ⎥
⎢  ∂ x3 ∂ x2 ⎥ −w23
i
an an n

not 1 ⎢ ∂ v 3 ∂ v1 ⎥
1 1
ω = rot (v) = ∇ × v ≡ ⎢ ⎥=⎢ ⎥
y ha

⎢ − − ⎥ ⎣ −w31 ⎦ (2.99)
2 2 2⎢ ∂ x1 ∂ x3 ⎥
le
⎢  ⎥ −w12
liv or ec

⎣ ∂ v 1 ∂ v2 ⎦
− −
M

.A

∂ x2 ∂ x1
m

ω = ∇ × v is named vorticiy vector


is extracted from (2.94). Vector 2ω 23 . It can
uu
e

be proven (in an analogous manner to Remark 2.20) that the equality


X Th

er
tin

ω ×r = w·r ∀r (2.100)
on

.O

is satisfied. Therefore, it is possible to characterize ω as the angular velocity of a


C

rotation motion, and ω × r = w · r as the rotation velocity of the point that has r
©

as the position vector with respect to the rotation center (see Figure 2.21). Then,
considering (2.90) and (2.92),
dv = l · dx = (d + w) · dx = d · dx + w · dx , (2.101)
     
stretch rotation
velocity velocity

which allows describing the relative velocity dv of the particles in the neigh-
borhood of a given particle P (see Figure 2.22) as the sum of a relative stretch
23 Observe the similarity in the structure of tensors Ω and θ in Section 2.11.6 and of tensors
w and ω seen here.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Material Time Derivatives of Strain and Other Magnitude Tensors 85

rs
ee
Figure 2.21: Vorticity vector.

s gin
t d le En

r
ba
ge ro or
eS m
ci
f

ra
C d P cs
b
a
i
an an n
y ha

le
liv or ec
M

.A

Figure 2.22: Stretch and rotation velocities.


m

d
uu
e

velocity (characterized by the strain rate tensor d) and a relative rotation velocity
X Th

ω ).
er

(characterized by the spin tensor w or the vorticity vector 2ω


tin
on

.O

2.14 Material Time Derivatives of Strain and Other


C

Magnitude Tensors
©

2.14.1 Deformation Gradient Tensor and its Inverse Tensor


Differentiating the expression of F in (2.3) with respect to time24 ,
∂ xi (X,t) dFi j ∂ ∂ xi (X,t) ∂ ∂ xi (X,t)
Fi j = =⇒ = = =⇒ (2.102)
∂ Xj dt ∂t ∂ X j ∂ Xj ∂t
  
vi

24 The Schwartz Theorem (equality of mixed partial derivatives) guarantees that


for a function Φ (x1 , x2 ... xn ) that is continuous and has continuous derivatives,
∂ 2 Φ/(∂ xi ∂ x j ) = ∂ 2 Φ/(∂ x j ∂ xi ) ∀i, j is satisfied.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
86 C HAPTER 2. S TRAIN

dFi j ∂ vi (X,t) ∂ vi (x (X,t)) ∂ xk


= = = lik Fk j =⇒
dt ∂ Xj ∂ xk ∂X
   j
lik Fk j

dF not .
= F = l ·F
dt (2.102 (cont.))
dFi j .
= Fi j = lik Fk j i, j ∈ {1, 2, 3}

rs
dt

ee
s gin
where (2.91) has been taken into account for the velocity gradient tensor l . To
obtain the material time derivative of tensor F−1 , the time derivative of the iden-
tity F · F−1 = 1 is performed25 .

t d le En
 

r
d   dF −1 d F−1

ba
ge ro or
−1 −1
F · F = 1 =⇒ F·F = ·F +F· =0

eS m
ci
 −1  dt dt dt
f

ra
d F .
C d P cs
=⇒ = −F−1 · F · F−1 = −F−1 · l · F · F−1 = −F−1 · l =⇒
b
a
dt    
i
an an n

l ·F 1
y ha

 −1 
le
liv or ec

d F
= −F−1 · l
M

dt
.A

(2.103)
dFi−1
= Fik−1 lk j
j
m

i, j ∈ {1, 2, 3}
d

dt
uu
e
X Th

er
tin

2.14.2 Material and Spatial Strain Tensors


on

.O

From (2.21), (2.102) and (2.93), it follows26


C

1 T  . 1 .T .
©

dE
E= F · F − 1 =⇒ =E= F · F + FT · F =
2 dt 2
1 T T  1  
= F · l · F + FT · l · F = FT · l + l T · F = FT · d · F
2 2   
.
=⇒ E = F · d · F .
T 2d (2.104)
 
25 The material time derivative of the inverse tensor d F−1 /dt must not be confused with
 . −1
the inverse of the material derivative of the tensor: F . These two tensors are completely
different tensors.
26 Observe that the result is the same as the one obtained in (2.98) using an alternative pro-
cedure.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Material Time Derivatives of Strain and Other Magnitude Tensors 87

Using (2.23) and (2.103) for the spatial strain tensor e yields

1 −T −1
 de . 1 d  −T  −1 d  −1 
e= 1−F ·F ⇒ =e=− F · F + F−T · F =
2 dt 2 dt dt
1  T −T −1 
= l · F · F + F−T · F−1 · l
2
. 1  T −T −1 
=⇒ e = l · F · F + F−T · F−1 · l . (2.105)
2

rs
2.14.3 Volume and Area Differentials

ee
The volume differential dV (X,t) associated with a certain particle P varies

s gin
along time (see Figure 2.23) and, in consequence, it makes sense to calculate

t d le En
its material derivative. Differentiating (2.55) for a volume differential results in

r
d d |F|

ba
ge ro or
dV (X,t) = |F (X,t)| dV0 (X) =⇒ dV (t) = dV0 .

eS m
(2.106)

ci
dt dt
f

ra
C d P cs
Therefore, the material derivative of the determinant of the deformation gradient
b
a
tensor |F| is27
i
an an n
y ha

d |F| d |F| dFi j dFi j


= |F| Fji−1 = |F| Fji−1 lik Fk j = |F| Fk j Fji−1 lik =
le
=
liv or ec

dt dFi j dt dt
   
M

.A

lik Fk j [ F·F−1 ] =δ
ki ki
m

∂ vi d |F|
uu

= |F| δki lik = |F| lii = |F| = |F| ∇ · v =⇒ = |F| ∇ · v , (2.107)


e

∂ xi dt
X Th

er
tin

where (2.102) and (2.91) have been considered. Introducing (2.107) into (2.106)
on

.O

and taking into account (2.55) finally results in


C

d
©

(dV ) = (∇ · v) |F| dV0 = (∇ · v) dV . (2.108)


dt
Operating in a similar manner yields the material derivative of the area dif-
ferential associated with a certain particle P and a given direction n (see Fig-
ure 2.24). The area differential vector associated with a particle in the reference
configuration, dA (X) = dA N, and in the present configuration, da (x,t) = da n,
are related through da = |F| · dA · F−1 (see (2.59)) and, differentiating this ex-

27 The derivative of the determinant of a tensor A with respect to the same tensor can be
written in compact notation as d |A|/dA = |A| · A−T or, in index notation, as d |A|/dAi j =
|A| · A−1
ji .

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
88 C HAPTER 2. S TRAIN

rs
Figure 2.23: Variation of the volume differential.

ee
s gin
pression, results in

t d le En
d d   d |F| d  −1 
(da) = |F| · dA · F−1 = dA · F−1 + |F| · dA =

r
F

ba
ge ro or
dt dt dt 
  dt  

eS m
ci
|F| ∇ · v −F−1 · l
f

ra
C d P cs
b
a
= (∇ · v) |F| dA · F−1 − |F| dA · F−1 · l =⇒
i
     
an an n
y ha

da da
le
liv or ec

d
(da) = (∇ · v) da − da · l = da · ((∇ · v) 1 − l ) , (2.109)
M

.A

dt
m

where (2.103) and (2.107) have been considered.


uu
e
X Th

er
tin
on

.O
C

Figure 2.24: Variation of the area differential.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Motion and Strains in Cylindrical and Spherical Coordinates 89

2.15 Motion and Strains in Cylindrical and Spherical


Coordinates
The expressions and equations obtained in intrinsic or compact notation are in-
dependent of the coordinate system considered. However, the expressions of the
components depend on the coordinate system used. In addition to the Cartesian
coordinate system, which has been used in the previous sections, two orthogonal
curvilinear coordinate systems will be considered here: cylindrical coordinates
and spherical coordinates.

rs
Remark 2.23. An orthogonal curvilinear coordinate system (gener-

ee
ically referred to as {a, b, c}), is characterized by its physical unit

s gin
basis {êa , êb , êc } (
êa
=
êb
=
êc
= 1), whose components are
orthogonal to each other (êa · êb = êa · êc = êb · êc = 0), as is also the

t d le En
case in a Cartesian system. The fundamental difference is that the

r
orientation of the curvilinear basis changes at each point in space

ba
ge ro or
(êm ≡ êm (x) m ∈ {a, b, c}). Therefore, for the purposes here, an

eS m
ci
f
orthogonal curvilinear coordinate system can be considered as a mo-

ra
bile Cartesian coordinate system {x , y , z } associated with a curvi-
C d P cs
b
a
linear basis {êa , êb , êc } (see Figure 2.25).
i
an an n
y ha

le
liv or ec
M

.A

Remark 2.24. The components, of a certain magnitude of vectorial


m

character (v) or tensorial character (T) in an orthogonal curvilinear


d

coordinate system {a, b, c}, can be obtained as the corresponding


uu
e

components in the local Cartesian system {x , y , z }:


X Th

er
tin

⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤
va vx  Taa Tab Tac Tx x Tx y Tx z
on

.O

not ⎢ ⎥ ⎢ ⎥ not ⎢ ⎥ ⎢ ⎥
v ≡ ⎣ vb ⎦ ≡ ⎣ vy ⎦ T ≡ ⎣ Tba Tbb Tbc ⎦ ≡ ⎣ Ty x Ty y Ty z ⎦
C

vc vz Tca Tcb Tcc Tz x Tz y Tz z

Remark 2.25. The curvilinear components of the differential opera-


tors (the ∇ operator and its derivatives) are not the same as their
counterparts in the local coordinate system {x , y , z }. They must be
defined specifically for each case. Their value for cylindrical and
spherical coordinates is provided in the corresponding section.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
90 C HAPTER 2. S TRAIN

2.15.1 Cylindrical Coordinates


The position of a certain point in space can be defined by its cylindrical coordi-
nates {r, θ , z} (see Figure 2.25). The figure also shows the physical orthonormal
basis {êr , êθ , êz }. This basis changes at each point in space according to
∂ êr ∂ êθ
= êθ and = −êr . (2.110)
∂θ ∂θ
Figure 2.26 shows the corresponding differential element. The expressions in
cylindrical coordinates of some of the elements treated in this chapter are:

rs
• Nabla operator, ∇

ee
 T

s gin
∂ 1 ∂ ∂ ∂
not 1 ∂ ∂
∇ = êr + êθ + êz =⇒ ∇≡ , , (2.111)
∂r r ∂θ ∂z ∂r r ∂θ ∂z

t d le En

r
ba
ge ro or
⎡ ⎤

eS m
x = r cos θ

ci
f x (r, θ , z) ≡ ⎣ y = r sin θ ⎦
not

ra
C d P cs
b z=z

a
i
an an n
y ha

le
liv or ec
M

.A
m

d
uu
e
X Th

er
tin

Figure 2.25: Cylindrical coordinates.


on

.O
C

Figure 2.26: Differential element in cylindrical coordinates.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Motion and Strains in Cylindrical and Spherical Coordinates 91

• Displacement vector, u, and velocity vector, v

u ≡ [ur , uθ , uz ]T
not
u = ur êr + uθ êθ + uz êz =⇒ (2.112)

v ≡ [vr , vθ , vz ]T
not
v = vr êr + vθ êθ + vz êz =⇒ (2.113)

• Infinitesimal strain tensor, ε


⎡ ⎤ ⎡ ⎤
εx x εx y εx z εrr εrθ εrz
1 

rs
not ⎢ ⎥ ⎢ ⎥
ε= (u ⊗ ∇) + (u ⊗ ∇)T ≡ ⎣ εx y εy y εy z ⎦ = ⎣ εrθ εθ θ εθ z ⎦

ee
2
εx z εy z εz z εrz εθ z εzz

s gin
∂ ur 1 ∂ uθ ur ∂ uz

t d le En
εrr = εθ θ = + εzz =
∂r r ∂θ r ∂z

r
 

ba
ge ro or
1 1 ∂ ur ∂ uθ uθ ∂ ur ∂ uz

eS m
1
εrθ = + − εrz = +

ci
2 r ∂θ ∂r f ∂z ∂r

ra
r 2
C d P cs

b
a
1 ∂ u θ 1 ∂ uz
i
an an n

εθ z = + (2.114)
2 ∂z r ∂θ
y ha

le
liv or ec

The components of ε are presented on the corresponding differential element in


M

.A

Figure (2.26).
m

• Strain rate tensor, d


uu
e
X Th

⎡ ⎤ ⎡ ⎤
er
tin

dx x dx y dx z drr drθ drz


1 
not ⎢ ⎥ ⎢ ⎥
d= (v ⊗ ∇) + (v ⊗ ∇)T ≡ ⎣ dx y dy y dy z ⎦ = ⎣ drθ dθ θ dθ z ⎦
on

.O

2
C

dx z dy z dz z drz dθ z dzz


©

∂ vr 1 ∂ vθ vr ∂ vz
drr = dθ θ = + dzz =
∂r r ∂θ r ∂z
 
1 1 ∂ vr ∂ vθ vθ 1 ∂ v r ∂ vz
drθ = + − drz = +
2 r ∂θ ∂r r 2 ∂z ∂r

1 ∂ vθ 1 ∂ vz
dθ z = + (2.115)
2 ∂z r ∂θ

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
92 C HAPTER 2. S TRAIN

2.15.2 Spherical Coordinates


A point in space is)defined by * its spherical coordinates {r, θ , φ }. The physical
orthonormal basis êr , êθ , êφ is presented in Figure 2.27. This basis changes at
each point in space according to
∂ êr ∂ êθ ∂ êφ
= êθ , = −êr and =0. (2.116)
∂θ ∂θ ∂θ
The expressions in spherical coordinates of some of the elements treated in this
chapter are:

rs
• Nabla operator, ∇

ee
 

s gin
∂ 1 ∂ 1 ∂ ∂not 1 ∂ 1 ∂ T
∇ = êr + êθ + êφ =⇒ ∇≡ , ,
∂r r ∂θ r sin θ ∂ φ ∂ r r ∂ θ r sin θ ∂ φ

t d le En
(2.117)

r
ba
ge ro or
eS m
ci
• Displacement vector, u, and velocity vector, v f

ra
C d P cs
b
a
not  T
i
u = ur êr + uθ êθ + uφ êφ =⇒ u ≡ ur , uθ , uφ
an an n

(2.118)
y ha

not  T
le
v = vr êr + vθ êθ + vφ êφ =⇒ v ≡ vr , vθ , vφ (2.119)
liv or ec
M

.A
m

d
uu
e
X Th

er
tin

⎡ ⎤
x = r sin θ cos φ
on

.O

x (r, θ , φ ) ≡ ⎣ y = r sin θ sin φ ⎦


not
C

z = z cos θ
©

Figure 2.27: Spherical coordinates.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Motion and Strains in Cylindrical and Spherical Coordinates 93

• Infinitesimal strain tensor, ε


⎡ ⎤ ⎡ ⎤
  εx  x  ε x  y  ε x  z  εrr εrθ εrφ
1 not ⎢ ⎥ ⎢ ⎥
ε= (u ⊗ ∇) + (u ⊗ ∇)T ≡ ⎣ εx y εy y εy z ⎦ = ⎣ εrθ εθ θ εθ φ ⎦
2
εx z εy z εz z εrφ εθ φ εφ φ

∂ ur 1 ∂ uθ ur
εrr = εθ θ = +
∂r r ∂θ r
1 ∂ uφ uθ ur

rs
εφ φ = + cot φ +
r sin θ ∂ φ r r

ee
 
1 1 ∂ ur ∂ uθ uθ 1 ∂ ur ∂ uφ uφ

s gin
1
εrθ = + − εrφ = + −
2 r ∂θ ∂r r 2 r sin θ ∂ φ ∂r r

t d le En

1 1 ∂ uθ 1 ∂ uφ uφ

r
εθ φ = + − cot φ (2.120)

ba
ge ro or
2 r sin θ ∂ φ r ∂θ r

eS m
ci
f

ra
The components of ε are presented on the corresponding differential element in
C d P cs
b
a
Figure 2.28.
i
an an n
y ha

le
• Strain rate tensor, d
liv or ec

⎡ ⎤ ⎡ ⎤
M

.A

  dx x dx y dx z drr drθ drφ


1 not ⎢ ⎥ ⎢ ⎥
d= (v ⊗ ∇) + (v ⊗ ∇)T ≡ ⎣ dx y dy y dy z ⎦ = ⎣ drθ dθ θ dθ φ ⎦
m

2
uu

dx z dy z dz z drφ dθ φ dφ φ


e
X Th

er
tin

∂ vr 1 ∂ vθ vr
drr = dθ θ = +
on

∂r r ∂θ
.O

r
1 ∂ vφ vθ
C

vr
dφ φ = + cot φ +
©

r sin θ ∂ φ r r
 
1 1 ∂ vr ∂ vθ vθ 1 1 ∂ vr ∂ vφ vφ
drθ = + − drφ = + −
2 r ∂θ ∂r r 2 r sin θ ∂ φ ∂r r

1 1 ∂ vθ 1 ∂ vφ vφ
dθ φ = + − cot φ (2.121)
2 r sin θ ∂ φ r ∂θ r

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
94 C HAPTER 2. S TRAIN

rs
ee
s gin
t d le En

r
ba
ge ro or
Figure 2.28: Differential element in spherical coordinates.

eS m
ci
f

ra
C d P cs
b
a
i
an an n
y ha

le
liv or ec
M

.A
m

d
uu
e
X Th

er
tin
on

.O
C

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Problems and Exercises 95

P ROBLEMS

Problem 2.1 – A deformation that takes place in a continuous medium has the
following consequences on the triangle shown in the figure below:
1. The segment OA increases its initial length in (1 + p).
2. The angle AOB decreases in q radians its initial value.

rs
3. The area increases its initial value in (1 + r).

ee
4. p, q, r, s 1.

s gin
The deformation is uniform and the z-axis is one of the principal directions of

t d le En
the deformation gradient tensor, which is symmetric. In addition, the stretch in
this direction is known to be λz = 1 + s. Obtain the infinitesimal strain tensor.

r
ba
ge ro or
eS m
ci
f

ra
C d P cs
b
a
i
an an n
y ha

le
liv or ec
M

.A
m

d
uu
e
X Th

er
tin
on

.O

Solution
C

A uniform deformation implies that the deformation gradient tensor (F) does
©

not depend on the spatial variables. Consequently, the strain tensor (E) and the
stretches (λ ) do not depend on them either. Also, note that the problem is to be
solved under infinitesimal strain theory.
The initial and final lengths of a segment parallel to the x-axis are related as
follows.
# A # A ⎫
OA f inal = λx dX = λx dX = λx OAinitial ⎬
O O =⇒ λx = 1 + p

OA f inal = (1 + p) OAinitial

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
96 C HAPTER 2. S TRAIN

Also, an initial right angle (the angle between the x- and y-axes) is related to its
corresponding final angle after the deformation through

π ⎪

initial angle = ⎬ q
2 =⇒ Δ Φ xy = −γxy = −2ε xy = −q =⇒ εxy = .
π ⎪
⎪ 2
final angle = + Δ Φ xy ⎭
2
In addition, F is symmetric and the z-axis is a principal direction, therefore
⎡ ⎤
∂ ux ∂ ux ∂ ux

rs
⎡ ⎤ ⎢1+ ∂x ∂y ∂z ⎥
⎢ ⎥

ee
F11 F12 0
not ⎢ ⎥ not not ⎢ ∂ uy ∂ uy ∂ uy ⎥
F ≡ ⎣ F12 F22 0 ⎦ ≡ 1 + J ≡ ⎢ ⎢ ⎥,

s gin
1+ ⎥
⎢ ∂ x ∂ y ∂ z ⎥
0 0 F33 ⎣ ∂ uz ∂ uz ∂ uz ⎦

t d le En
1+
∂x ∂y ∂z

r
ba
ge ro or
eS m
which reveals the nature of the components of the displacement vector,

ci
⎧  f

ra
C d P cs
⎪ ∂ ∂
bux (x, y) ,

a

⎪ u x u y
⎨ = = 0 =⇒
i
an an n

∂z ∂z uy (x, y) ,
y ha


⎪ ∂ u z ∂ uz

⎩ = = 0 =⇒ uz (z) .
le
liv or ec

∂x ∂y
M

.A

Then, the following components of the strain tensor can be computed.


m


d

1 ∂ ux ∂ u z
uu

εxz = + = 0 =⇒ εxz = 0
e

2 ∂z ∂x
X Th

er


tin

1 ∂ ux ∂ u z
εxz = + = 0 =⇒ εxz = 0
on

.O

2 ∂z ∂x ⎫
∂ uz ⎬
C

εzz = = λz − 1
∂z =⇒ εzz = s
©

λ = 1+s ⎭
z

In infinitesimal strain theory, F = 1 + ε + Ω , where Ω33 = 0 since the infinites-


imal rotation tensor is antisymmetric. Thus, Fzz = 1 + εzz results in Fzz = 1 + s .
Now, the relation between the initial and final areas is dA = |F| dA0 ·F−1 , where
the inverse tensor of F is calculated using the notation
⎡ ⎤
B11 B12 0 + ,
not ⎢ ⎥ C C
F ≡ ⎣ B12 B22 0 ⎦ with B−1 ≡
not 11 12
,
C12 C22
0 0 1+s

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Problems and Exercises 97

which yields the inverse tensor of F,


⎡ ⎤
C11 C12 0
⎢ ⎥
F−1 ≡ ⎢ 0 ⎥.
not
C C
⎣ 12 22 ⎦
1
0 0
1+s
The area differential vector is defined as
⎡ ⎤ ⎡ ⎤
0 0

rs
not ⎢ ⎥ not ⎢ ⎥
dA0 ≡ ⎣ 0 ⎦ =⇒ dA0 · F−1 ≡ ⎢ ⎣
0 ⎥.

ee
1
dA0 dA0

s gin
1+s

t d le En
Then, taking into account that |F| = Tr (εε ) + 1, and neglecting second-order
terms results in

r
ba

ge ro or
eS m
dA = (1 + r) dA0 ⎬

ci
f
=⇒ εyy = r − p .

ra
1
C d P cs
dA = (1 + p + s + εyy ) dA0 ⎭
b
a
1+s
i
an an n
y ha

Finally, since the strain tensor is symmetric,


le
⎡ ⎤
liv or ec

q
M

.A

⎢p 2
0⎥
not ⎢ q ⎥
ε ≡⎢ 0⎥⎥ .
m

⎢2 r− p
d

⎣ ⎦
uu
e

0 0 s
X Th

er
tin
on

.O
C

Problem 2.2 – A uniform deformation (F = F (t)) is produced on the tetrahe-


©

dron shown in the figure below, with the following consequences:

1. Points O, A and B do not move.


2. The volume of the solid becomes p times
its initial volume. √
3. The length of segment AC becomes p/ 2
times its initial length.
4. The final angle AOC has a value of 45◦ .

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
98 C HAPTER 2. S TRAIN

Then,
a) Justify why the infinitesimal strain theory cannot be used here.
b) Determine the deformation gradient tensor, the possible values of p and the
displacement field in its material and spatial forms.
c) Draw the deformed solid.

Solution

rs
a) The angle AOC changes from 90◦ to 45◦ therefore, it is obvious that the

ee
deformation involved is not infinitesimal. In addition, under infinitesimal strain
theory Δ Φ 1 is satisfied and, in this problem, Δ Φ = π/4 ≈ 0.7854.

s gin
Observation: strains are dimensionless; in engineering, small strains are usually

t d le En
considered when these are of order 10−3 − 10−4 .

r
ba
ge ro or
eS m
b) The conditions in the statement of the problem must be imposed one by one:

ci
f
1. Considering that F (X,t) = F (t) and knowing that dx = F · dX, the latter

ra
C d P cs
b
can be integrated as

a
i
# # #
an an n
y ha

x= dx = FdX = F dX = F (t) · X + C (t)


le
liv or ec

⎡ ⎤ ⎡ ⎤
F11 F12 F13 C1
M

.A

not ⎢ ⎥ not ⎢ ⎥
with F ≡ ⎣ F21 F22 F23 ⎦ and C ≡ ⎣ C2 ⎦ ,
m

F31 F32 F33 C3


uu
e
X Th

er

which results in 12 unknowns. Imposing now the conditions in the statement,


tin

Point O does not move:


on

.O

⎡ ⎤ ⎡ ⎤ ⎡ ⎤
0 0 0
C

⎢ ⎥ ⎢ ⎥ not ⎢ ⎥
⎣ 0 ⎦ = [F] ⎣ 0 ⎦ + C =⇒ C ≡ ⎣ 0 ⎦
©

0 0 0

Point A does not move:


⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎧
a a a F11 ⎨ F11 = 1

⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ 0 ⎦ = [F] ⎣ 0 ⎦ = ⎣ a F21 ⎦ =⇒ F21 = 0


0 0 a F31 F31 = 0

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Problems and Exercises 99

Point B does not move:


⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎧
0 0 a F12 ⎨ F12 = 0

⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎣ a ⎦ = [F] ⎣ a ⎦ = ⎣ a F22 ⎦ =⇒ F22 = 1


0 0 a F32 F32 = 0

Grouping all the information obtained results in


⎡ ⎤
1 0 F13
not ⎢ ⎥
F ≡ ⎣ 0 1 F23 ⎦ .

rs
ee
0 0 F33

s gin
2. The condition in the statement imposes that V f inal = pVinitial .

t d le En
Expression dV f = |F| dV0 allows to locally relate the differential volumes at
different instants of time. In this case, F is constant for each fixed t, thus, the

r
ba
ge ro or
expression can be integrated and the determinant of F can be moved outside the

eS m
ci
integral, # # #
f

ra
C d P cs
Vf = dV f = |F| dV0 = |F| dV0 = |F| V0 .
b
a
i
V V0 V0
an an n

Therefore, |F| = F33 = p must be imposed.


y ha

le
liv or ec

3. The condition in the statement imposes that lAC, f inal = √p lAC, initial .
2
M

.A

Since F is constant, the transformation is linear, that is, it transforms straight


lines into straight lines. Hence, AC in the deformed configuration must also be a
m

rectilinear segment. Then,


uu
e

⎡ ⎤⎡ ⎤ ⎡ ⎤
X Th

er
tin

1 0 F13 0 a F13
not ⎢ ⎥⎢ ⎥ ⎢ ⎥
xC = F · XC ≡ ⎣ 0 1 F23 ⎦ ⎣ 0 ⎦ = ⎣ a F23 ⎦ and
on

.O

0 0 F33 a ap
C

" " " "


lAC, f inal = lAC = " [a F13 , a F23 , ap] − [a, 0, 0] " = " [a (F13 − 1) , a F23 , ap] " =
! !
= (a (F13 − 1))2 + (a F23 )2 + (ap)2 = a (F13 − 1)2 + F23 2 + p2 =

p p √
= √ lAC = √ 2 a = p a .
2 2
Therefore,
!
(F13 − 1)2 + F23
2 + p2 = p ⇒ (F − 1)2 + F 2 = 0 ⇒ F = 1; F = 0
13 23 13 23

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
100 C HAPTER 2. S TRAIN

and the deformation gradient tensor results in


⎡ ⎤
1 0 1
not ⎢ ⎥
F ≡ ⎣0 1 0⎦ ,
0 0 p

such that only the value of p remains to be found.

4. The condition in the statement imposes that AOC f inal = 45◦ = π/4.

rs
Considering dX(1) ≡ [1, 0, 0] and dX(2) ≡ [0, 0, 1], the corresponding vectors
not not

ee
in the spatial configuration are computed as

s gin
⎡ ⎤⎡ ⎤ ⎡ ⎤
1 0 1 1 1
(1) not ⎢ ⎥⎢ ⎥ ⎢ ⎥

t d le En
(1)
dx = F · dX ≡ ⎣ 0 1 0 ⎦ ⎣ 0 ⎦ = ⎣ 0 ⎦ ,

r
0 0 p 0 0

ba
ge ro or
eS m
ci

⎤⎡ ⎤ ⎡ ⎤
f

ra
0 1 0 1
1
C d P cs
(2) not ⎢ ⎥⎢ ⎥ ⎢ ⎥
b
a
(2)
dx = F · dX ≡ ⎣ 0 1 0 ⎦ ⎣ 0 ⎦ = ⎣ 0 ⎦ .
i
an an n
y ha

0 0 p 1 p
le
liv or ec

Then, √
  dx(1) · dx(2) 2
M

.A


cos AOC f inal = cos 45 = "" (1) "" "" (2) "" =
dx dx 2
m

is imposed, with
uu
e

" " " "


" (2) " $
X Th

" (1) "


er
tin

"dx " = 1 , "dx " = 1 + p2 and dx(1) · dx(2) = 1


on

.O

such that √
C

1 2 1
$ = =√ =⇒ p = ±1 .
©

1 + p2 2 2
But |F| = p > 0, and, consequently, p = 1. Then, the deformation gradient tensor
is
⎡ ⎤
1 0 1
not ⎢ ⎥
F ≡ ⎣0 1 0⎦ .
0 0 1

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Problems and Exercises 101

The equation of motion is determined by means of x = F · X,


⎡ ⎤ ⎡ ⎤⎡ ⎤ ⎡ ⎤
x 1 0 1 X X +Z
⎢ ⎥ ⎢ ⎥⎢ ⎥ ⎢ ⎥
⎣y⎦ = ⎣0 1 0⎦⎣Y ⎦ = ⎣ Y ⎦ ,
z 0 0 1 Z Z

which allows determining the displacement field in material and spatial descrip-
tions as
⎡ ⎤ ⎡ ⎤

rs
Z z
not ⎢ ⎥ not ⎢ ⎥

ee
U (X,t) = x − X ≡ ⎣ 0 ⎦ and u (x,t) ≡ ⎣ 0 ⎦ .

s gin
0 0

t d le En
c) The graphical representation of the deformed tetrahedron is:

r
ba
ge ro or
eS m
ci
f

ra
C d P cs
b
a
i
an an n
y ha

le
liv or ec
M

.A
m

d
uu
e
X Th

er
tin
on

.O

Problem 2.3 – A uniform deformation is applied on the solid shown in the


figure below. Determine:
C

a) The general expression of the material description of the displacement field


U (X,t) in terms of the material displacement gradient tensor J.
b) The expression of U (X,t) when, in addition, the following boundary condi-
tions are satisfied:

UY = UZ = 0 , ∀ X, Y, Z
"
UX "X=0 = 0 , ∀ X, Y
"
UX "X=L = δ

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
102 C HAPTER 2. S TRAIN

c) The possible values (positive and negative) that δ may take. Justify the an-
swer obtained.
d) The material and spatial strain tensors and the infinitesimal strain tensor.
e) Plot the curves EXX − δ /L, exx − δ /L and εx − δ /L for all possible values
of δ , indicating every significant value.

rs
ee
s gin
t d le En

r
ba
ge ro or
eS m
ci
f

ra
Solution
C d P cs
b
a
a) A uniform deformation implies that F (X,t) = F (t) , ∀t, X. The deformation
i
an an n

gradient tensor is related to the material displacement gradient tensor through


y ha

the expression F = 1 + J. Therefore, if F = F (t), then J = J (t). Taking into


le
account the definition of J and integrating its expression results in
liv or ec

# #
M

.A

∂ U (X,t)
J= =⇒ dU = J dX =⇒ dU = J dX
∂X
m

# #
uu
e

=⇒ dU = J dX =⇒ U = J · X + C (t) .
X Th

er
tin

where C (t) is an integration constant. Then, the general expression of the mate-
on

.O

rial description of the displacement field is


C

U (X,t) = J (t) · X + C (t) .

b) Using the previous result and applying the boundary conditions given in the
statement of the problem will yield the values of J and C.
Boundary conditions:
UY = UZ = 0 , ∀ X, Y, Z ⇒ Points only move in the X-direction.
"
UX "X=0 = 0 , ∀ Y, Z ⇒ The YZ plane at the origin is fixed.
"
UX "X=L = δ , ∀ Y, Z ⇒ This plane moves in a uniform manner
in the X-direction.
X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Problems and Exercises 103

If the result obtained in a) is written in component form, the equations and con-
clusions that can be reached will be understood better.

UX = J11 X + J12 Y + J13 Z +C1


UY = J21 X + J22 Y + J23 Z +C2
UZ = J31 X + J32 Y + J33 Z +C3

From the first boundary condition:

UY = 0 , ∀ X, Y, Z =⇒ J21 = J22 = J23 = C2 = 0

rs
UZ = 0 , ∀ X, Y, Z =⇒ J31 = J32 = J33 = C3 = 0

ee
s gin
From the second boundary condition:
"

t d le En
UX "X=0 = 0 , ∀ Y, Z =⇒ J12 = J13 = C1 = 0

r
ba
ge ro or
From the third boundary condition:

eS m
ci
" f

ra
δ
C d P cs
UX "X=L = δ , ∀ Y, Z =⇒ J11 L = δ ⇒ J11 =
b
a
L
i
an an n
y ha

Finally,
⎡ ⎤ ⎡ ⎤
⎡ ⎤ δ
le
δ
liv or ec

0 0 0 ⎢L ⎥X
⎢ L ⎥ not ⎢ ⎥ not ⎢ ⎥
M

.A

J≡⎢ ⎥
not
⎣0 0 0⎦; C ≡ ⎣0⎦ =⇒ U (X) = J · X + C ≡ ⎢ 0 ⎥ .
⎣ ⎦
m

0 0 0 0
d

0
uu
e
X Th

er
tin

c) In order to justify all the possible positive and negative values that δ may
on

.O

take, the condition |F| > 0 must be imposed. Therefore, the determinant of F
must be computed,
C

⎡ ⎤
δ
1+ 0 0
not ⎢ ⎥
F = 1+J ≡ ⎢
L ⎥ =⇒ |F| = 1 + δ > 0 =⇒ δ > −L .
⎣ 0 1 0 ⎦ L
0 0 1

d) To obtain the spatial and material strain tensors as well as the infinitesimal
strain tensor, their respective definitions must be taken into account.
1 
Spatial strain tensor: e= 1 − F−T · F−1
2

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
104 C HAPTER 2. S TRAIN

1 T 
Material strain tensor: E=
F ·F−1
2
1 T 
Infinitesimal strain tensor: ε = J ·J
2
Applying these definitions using the values of F and J calculated in b) and c),
the corresponding expressions are obtained.

⎡ ⎤
exx 0 0  -
⎢ ⎥ δ 1 δ2 δ 2

rs
not
e ≡ ⎣ 0 0 0 ⎦ with exx = + 1 +
L 2 L2 L

ee
0 0 0
⎡ ⎤ ⎡ ⎤

s gin
δ
EXX 0 0 ⎢L 0 0 ⎥
not ⎢ ⎥
⎥ with EXX = δ + 1 δ
2
not ⎢ ⎥

t d le En
E≡⎢ ⎣ 0 0 0 ⎦ ; ε ≡ ⎢ 0 0 0 ⎥
L 2L 2 ⎣ ⎦

r
ba
ge ro or
0 0 0 0 0 0

eS m
ci
f

ra
C d P cs
b
a
i
an an n

e) Plotting the curves EXX − δ /L , exx − δ /L and εx − δ /L together yields:


y ha

le
liv or ec

Here,
M

.A

• EXX is a second-order
parabola that contains the
m

origin and has its mini-


uu
e

mum at δ /L = −1, i.e., for


X Th

er

EXX = −1/2.
tin

• εx is the identity straight line


on

.O

(45◦ slope and contains the


C

origin).
©

• exx has two asymptotes, a


vertical one at δ /L = −1 and
a horizontal at exx = 1/2.

It can be concluded, then, that for small δ /L strains the three functions have a
very similar behavior and the same slope at the origin. That is, the same result
will be obtained with any of the definitions of strain tensor. However, outside
this domain (large or finite strains) the three curves are clearly different.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Problems and Exercises 105

E XERCISES

2.1 – Consider the velocity fields


   T
not x 2y 3z T X
not 2Y 3Z
v1 ≡ , , and v2 ≡ , , .
1+t 1+t 1+t 1+t 1+t 1+t
Determine:

rs
a) The material description of v1 and the spatial description of v2 (consider

ee
t = 0 is the reference configuration).

s gin
b) The density distribution in both cases (consider ρ0 is the initial density).
c) The material and spatial descriptions of the displacement field as well as

t d le En
the material (Green-Lagrange) and spatial (Almansi) strain tensors for the

r
velocity field v1 .

ba
ge ro or
eS m
d) Repeat c) for configurations close to the reference configuration (t → 0).

ci
f

ra
C d P cs
e) Prove that the two strain tensors coincide for the conditions stated in d).
b
a
i
an an n
y ha

2.2 – The equation of motion in a continuous medium is


le
liv or ec

x = X +Y t , y=Y , z=Z.
M

.A

Obtain the length at time t = 2 of the segment of material line that at time t = 1
m

is defined in parametric form as


d
uu
e

x (α) = 0 , y (α) = α 2 , z (α) = α 0≤α ≤1.


X Th

er
tin
on

.O

2.3 – Consider the material strain tensor


C

⎡ ⎤
©

0 tetX 0
⎢ ⎥
E≡⎢ 0 ⎥
not
⎦.
tX
⎣ te 0
0 0 tetY

Obtain the length at time t = 1 of the segment that at time t = 0 (reference


configuration) is straight and joins the points (1, 1, 1) and (2, 2, 2).

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
106 C HAPTER 2. S TRAIN

2.4 – The equation of motion of a continuous medium is

x=X , y=Y , z = Z − Xt .
Calculate the angle formed at time t = 0 by the differential segments that at time
t = t are parallel to the x- and z-axes.

2.5 – The following information is known in relation to a certain displacement


field given in material description, U (X,Y, Z):
1) It is lineal in X, Y , Z.

rs
2) It is antisymmetric with re-

ee
spect to plane Y = 0, that is, the
following is satisfied:

s gin
U (X,Y, Z) = −U (X, −Y, Z)

t d le En
∀ X,Y, Z

r
ba
ge ro or
eS m
ci
3) Under said displacement field,
f

ra
C d P cs
the volume of the element in the
b
a
figure does not change, its an-
i
an an n

gle AOB remains constant,√ the


y ha

segment OB becomes 2 times


le
liv or ec

its initial length and the vertical


component of the displacement at
M

.A

point B is positive (wB > 0).


m

Determine:
d
uu
e

a) The most general expression of the given displacement field, such that condi-
X Th

er
tin

tions 1) and 2) are satisfied.


b) The expression of U when, in addition, condition 3) is satisfied. Obtain the
on

.O

deformation gradient tensor and the material strain tensor. Draw the de-
C

formed shape of the element in the figure, indicating the most significant
©

values.
c) The directions (defined by their unit vectors T) for which the deformation is
reduced to a stretch (there is no rotation).
NOTE: Finite strains must be considered (not infinitesimal ones).

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
Problems and Exercises 107

2.6 – The solid in the figure undergoes a uniform deformation such that points
A, B and C do not move. Assuming an infinitesimal strain framework,
a) Express the displacement field in terms of “generic” values of the stretches
and rotations.
b) Identify the null components of the strain tensor and express the rotation
vector in terms of the stretches.

In addition, the following is known:


1) Segment AE becomes (1 + p)
times its initial length.

rs
2) The volume becomes (1 + q)

ee
times its initial value.

s gin
3) The angle θ increases its value
in r (given in radians).

t d le En
Under these conditions, deter-

r
ba
mine:

ge ro or
eS m
ci
c) The strain tensor, the rotation
f

ra
C d P cs
vector and the displacement
b
a
i
field in terms of p, q and r.
an an n
y ha

NOTE: The values of p, q and


le
liv or ec

r are small and its second-order


infinitesimal terms can be ne-
M

.A

glected.
m

d
uu

2.7 – The solid in the figure undergoes a uniform deformation with the following
e

consequences:
X Th

er
tin

1) The x- and z-axes are both


on

material lines. Point A does


.O

not move.
C

2) The volume of the solid re-


©

mains constant.
3) The angle θxy remains con-
stant.
4) The angle θyz increases in r
radians.
5) The segment AF becomes
(1 + p) times its initial length.
6) The area of the triangle
ABE becomes (1 + q) its ini-
tial value.

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961
108 C HAPTER 2. S TRAIN

Then,
a) Express the displacement field in terms of “generic” values of the stretches
and rotations.
b) Identify the null components of the strain tensor and express the rotation
vector in terms of the stretches.
c) Determine the strain tensor, the rotation vector and the displacement field in
terms of p, q and r.
NOTE: The values of p, q and r are small and its second-order infinitesimal
terms can be neglected.

rs
ee
2.8 – The sphere in the figure undergoes a uniform deformation (F = const.)

s gin
such that points A, B and C move to positions A , B and C , respectively. Point
O does not move. Determine:

t d le En
a) The deformation gradient tensor in terms of p and q.

r
ba
ge ro or
b) The equation of the deformed external surface of the sphere. Indicate which

eS m
ci
type of surface it is and draw it.
f

ra
C d P cs
c) The material and spatial strain tensors. Obtain the value of p in terms of q
b
a
i
when the material is assumed to be incompressible.
an an n
y ha

d) Repeat c) using infinitesimal strain theory. Prove that when p and q are small,
the results of c) and d) coincide.
le
liv or ec
M

.A
m

d
uu
e
X Th

er
tin
on

.O
C

X. Oliver and C. Agelet de Saracibar Continuum Mechanics for Engineers.Theory and Problems
doi:10.13140/RG.2.2.25821.20961

Вам также может понравиться