Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Tutorial 3

Analysis of strain and elastic stress-strain relationship

Course of Machine Design 2


Politecnico di Milano, Dipartimento di Meccanica

March 2019
Definition of strain

Let us imagine a bar that is stretched and let us consider a segment A − B. Because
of the stretching the two ends of the segment move along the bar to A0 − B 0 .

Deformation of a bar
Definition of strain

Let us imagine a bar that is stretched and let us consider a segment A − B. Because
of the stretching the two ends of the segment move along the bar to A0 − B 0 .
The segment has been stretched
of A0 B 0 − AB = ∆u.

Deformation of a bar
Definition of strain

Let us imagine a bar that is stretched and let us consider a segment A − B. Because
of the stretching the two ends of the segment move along the bar to A0 − B 0 .
The segment has been stretched
of A0 B 0 − AB = ∆u.
The strain is:
∆u du
x = lim = (1)
∆x→0 ∆x dx
(+) –> elongation
(-) –> contraction
Deformation of a bar
If the bar is uniformly stretched with a total displacement δ, the strain can be
calculated as:
L − Lo δ
o = =
Lo Lo
Definition of strain

Let us imagine a bar that is stretched and let us consider a segment A − B. Because
of the stretching the two ends of the segment move along the bar to A0 − B 0 .
The segment has been stretched
of A0 B 0 − AB = ∆u.
The strain is:
∆u du
x = lim = (1)
∆x→0 ∆x dx
(+) –> elongation
(-) –> contraction
Deformation of a bar
If the bar is uniformly stretched with a total displacement δ, the strain can be
calculated as:
L − Lo δ
o = =
Lo Lo
If the deformation is not uniform, this is a simple average strain.
Displacements and deformations in 2D - 1

Lets us imagine that in body a subjected to external loads we have displacements so


that all the points remain in the same plane, we want to calculate the corresponding
strains, under the hypothesis of small displacements.
Let us consider three points A-B-C which encompass an angle π/2 delimited by
segments AB and AC , whose initial length is dx and dy .

If we consider the segment AB:


u(x,y+dy,z)
∆l u(x + dx, y , z, t) − u(x, y , z, t)
x = =
l dx
C that can be written (small displacements):

dy ∆l ∂u
x = = (2)
l ∂x
In the deformed configuration, the segment AC
u(x,y,z) u(x+dx,y,z) rotates around z of an angle:
A B
dx
u(x, y + dy , z, t) − u(x, y , z, t) ∂u
Displacement u of the points A,B and = (3)
C after body deformation dy ∂y
Displacements and deformations in 2D - 2
Lets us consider the strain components due to the displacement along the y direction.

If we consider the segment AC :

v (x, y + dy , z, t) − v (x, y , z, t) ∂v
y = = (4)
C dy ∂y

v(x+dx,y,z) Similarly to the u displacements, the rotation of AB


dy
around the axis z is:
v(x,y,z) v (x + dx, y , z, t) − v (x, y , z, t) ∂v
= (5)
A B dx ∂x
dx
Displacement v of the points A,B and The shear displacement γxy (the total variation of
C ˆ ) in the plane xy is:
ABC

∂u ∂v
γxy = + . (6)
∂y ∂x
Displacements and deformations in 2D - 2
Lets us consider the strain components due to the displacement along the y direction.

If we consider the segment AC :

v (x, y + dy , z, t) − v (x, y , z, t) ∂v
y = = (4)
C dy ∂y

v(x+dx,y,z) Similarly to the u displacements, the rotation of AB


dy
around the axis z is:
v(x,y,z) v (x + dx, y , z, t) − v (x, y , z, t) ∂v
= (5)
A B dx ∂x
dx
Displacement v of the points A,B and The shear displacement γxy (the total variation of
C ˆ ) in the plane xy is:
ABC

∂u ∂v
γxy = + . (6)
∂y ∂x
Displacements and deformations in 2D - 2
Lets us consider the strain components due to the displacement along the y direction.

If we consider the segment AC :

v (x, y + dy , z, t) − v (x, y , z, t) ∂v
y = = (4)
C dy ∂y

v(x+dx,y,z) Similarly to the u displacements, the rotation of AB


dy
around the axis z is:
v(x,y,z) v (x + dx, y , z, t) − v (x, y , z, t) ∂v
= (5)
A B dx ∂x
dx
Displacement v of the points A,B and The shear displacement γxy (the total variation of
C ˆ ) in the plane xy is:
ABC

∂u ∂v
γxy = + . (6)
∂y ∂x

The strain components can be so viewed as (positive strains are displayed):


Strain tensor - 1

For a 3D solid the state of deformation at a point is defined by the six components:

∂u ∂v ∂w
x = , y = , z =
∂x ∂y ∂z
(7)
∂u ∂v ∂v ∂w ∂u ∂w
γxy = + , γyz = + , γxz = +
∂y ∂x ∂z ∂y ∂z ∂x

Eq. (7), if we write γxy = 2xy , can be replaced by the synthetic notation 1 :
 
1 ∂ui ∂uj
ij = + (8)
2 ∂xj ∂xi

1
we simplify: ii = i .
Strain tensor - 1

For a 3D solid the state of deformation at a point is defined by the six components:

∂u ∂v ∂w
x = , y = , z =
∂x ∂y ∂z
(7)
∂u ∂v ∂v ∂w ∂u ∂w
γxy = + , γyz = + , γxz = +
∂y ∂x ∂z ∂y ∂z ∂x

Eq. (7), if we write γxy = 2xy , can be replaced by the synthetic notation 1 :
 
1 ∂ui ∂uj
ij = + (8)
2 ∂xj ∂xi

ˆ , we could write γxy = γyx .


Since γxy is the variation or distortion of ABC

1
we simplify: ii = i .
Strain tensor - 1

For a 3D solid the state of deformation at a point is defined by the six components:

∂u ∂v ∂w
x = , y = , z =
∂x ∂y ∂z
(7)
∂u ∂v ∂v ∂w ∂u ∂w
γxy = + , γyz = + , γxz = +
∂y ∂x ∂z ∂y ∂z ∂x

Eq. (7), if we write γxy = 2xy , can be replaced by the synthetic notation 1 :
 
1 ∂ui ∂uj
ij = + (8)
2 ∂xj ∂xi

ˆ , we could write γxy = γyx .


Since γxy is the variation or distortion of ABC

On the top of the symmetry ij = ji , strains can be represented by a tensor:
 
xx xy xz
[ij ] = xy yy yz  (9)
xz yz zz

1
we simplify: ii = i .
Strain tensor - 2

If we know the strain tensor in a refernce X − Y − Z , the strains 0ij in a new reference
X 0 − Y 0 − Z 0 can be once again obtained with the same equation used for the stress
tensor:
[0ij ] = T · [ij ] · T T (10)
The tensor [ij ] has the same properties as the stress tensor, namely it exists a triplet
of directions which identify the planes of principal strains.
Strain tensor - 2

If we know the strain tensor in a refernce X − Y − Z , the strains 0ij in a new reference
X 0 − Y 0 − Z 0 can be once again obtained with the same equation used for the stress
tensor:
[0ij ] = T · [ij ] · T T (10)
The tensor [ij ] has the same properties as the stress tensor, namely it exists a triplet
of directions which identify the planes of principal strains.
The principal strains 1 − 2 − 3 can be found: i) by calculating with suitable
algorithms the eigenvalues of the matrix ij ; ii) by finding the values p which satisfy
the characteristic equation:

(xx − p ) xy xz


xy (yy − p ) yz =0
(11)
xz yz (zz − p )

By solving the determinant, it is possible to write a cubic equation:

3p − E1 2p + E2 p − E3 = 0 (12)

where E1 − E2 − E3 are called strain invariants (as we have already seen for the stress
tensors, they can simply expressed in terms of the eigenvalues).The p values are the
roots of this equation.
Strain in one direction in plane problems

In a 2D problem, once the strain tensor is Y' Y


known in a X − Y reference, then the
strain tensor in a X 0 − Y 0 system rotated
of an angle θ can be expressed by (10).
X'
 
cos θ sin θ θ
T =
− sin θ cos θ X
Strain in one direction in plane problems

In a 2D problem, once the strain tensor is Y' Y


known in a X − Y reference, then the
strain tensor in a X 0 − Y 0 system rotated
of an angle θ can be expressed by (10).
X'
 
cos θ sin θ θ
T =
− sin θ cos θ X

If we take the normal strain in the X 0 direction, at an angle θ from the axis X (the
angle is positive counterclockwise), we then have:

X 0 = x cos2 θ + y sin2 θ + γxy sin θ cos θ (13)


Strain in one direction in plane problems

In a 2D problem, once the strain tensor is Y' Y


known in a X − Y reference, then the
strain tensor in a X 0 − Y 0 system rotated
of an angle θ can be expressed by (10).
X'
 
cos θ sin θ θ
T =
− sin θ cos θ X

If we take the normal strain in the X 0 direction, at an angle θ from the axis X (the
angle is positive counterclockwise), we then have:

X 0 = x cos2 θ + y sin2 θ + γxy sin θ cos θ (13)

This expression is used for analysing strain gauge measurements in plane problems.
In particular, in the case that we do not know the principal directions, we will need
three measurements for determining the components of [ij ], since (if the principal
directions are not known) we have three unknown values (x , y , γxy ).
Strain gauge rosette

Strain measurements in plane problems: a) rosette with 3 grids at 120 deg; b) rosette
-45/0/45 deg; c) general schematic of a rosette.
Strain gauge rosette

Strain measurements in plane problems: a) rosette with 3 grids at 120 deg; b) rosette
-45/0/45 deg; c) general schematic of a rosette.

Special strain-gage arrangements are called strain-gage rosettes: they consist of three
grids usually located at 120 deg (Y configuration) or 0 − 45 − 90 degs.
If we consider a rosette with grids a − b − c at the angles θa − θb − θc respect to X
axis, the strains for the three sensors are:
 2 2
 a = x cos θa + y sin θa + γxy sin θa cos θa


b = x cos2 θb + y sin2 θb + γxy sin θb cos θb (14)

c = x cos2 θc + y sin2 θc + γxy sin θc cos θc

By solving the system of equations it is then possible to determine [ij ] and then
principal strains.
Volume change

C+dC Let us consider a material volume V = A × B × C ,


C with the faces parallel to X − Y − Z . If the three
sides have been stretched, the strains are:
y
A
dA dB dC
B A+dA x = , y = , z =
dx dy dz
B+dB
x
Calculation of volumetric strain v
Volume change

C+dC Let us consider a material volume V = A × B × C ,


C with the faces parallel to X − Y − Z . If the three
sides have been stretched, the strains are:
y
A
dA dB dC
B A+dA x = , y = , z =
dx dy dz
B+dB
x
Calculation of volumetric strain v

the change of material volume V could be expressed as:

∂V ∂V ∂V
dV = dA + dB + dC = BC · dA + AC · dB + AB · dC (15)
∂A ∂B ∂C
dividing by V we obtain the volumetric strain or dilatation:

dV dA dB dC
= V = + + = x + y + z (16)
V A B C
Compatibility equations - 1

We have obtained the relationships among strains and displacements: given 3


continuous functions for the diplacements u - v - w , it is then possible to obtain 6
strain components. It is then simple to imagine that, if we integrated the ij for
obtaining the displacements, the 6 strain components cannot be independent and
there should be some relationship between them.

Let us consider a solid discretized in


elements in (a) and let us observe the
undeformed configuration in (b). Let us
imagine how the elements can be
deformed: in (c) the deformed
configuration allows to have continuity
with the neighbouring elements so that a
continuous displacement field, while in (d)
the deformed elements do not match with
the adjacent ones.

A relationship between the strains


should exists so that a picture like (c)
could be obtained.
Compatibility equations - 2 plane problems

From Eq.7, for a 2D solid we can simply write:

∂u ∂v ∂u ∂v
x = , y = , γxy = + (17)
∂x ∂y ∂y ∂x

by successive derivations, it is possible then to write the relationships is:

∂ 2 x ∂3u ∂ 2 y ∂3v ∂ 2 γxy ∂3u ∂3v


= = = +
∂y 2 ∂y 2 ∂x ∂x 2 ∂x 2 ∂y ∂x∂y ∂y 2 ∂x ∂x 2 ∂y

from these ones we obtain the compatibility equations:

∂ 2 x ∂ 2 y ∂ 2 γxy
+ = (18)
∂y 2 ∂x 2 ∂x∂y

The condition expressed by differential equation has be fullfilled by the strain


components, so that two continuous functions for u e v could express the strains with
Eq. 17.
Compatibility equations- 3D case

In 3D problems, two other equations similar to Eq. 18 can be written by permutations


of the indexes x − y − z obtaining the system of equations
 2
∂ x ∂ 2 y ∂ 2 γxy
 ∂y 2 + ∂x 2 = ∂x∂y






 2
∂ y ∂ 2 z ∂ 2 γyz
+ = (19)
 ∂z 2
 ∂y 2 ∂y ∂z



∂ 2 ∂ 2 ∂ 2γ
x z xz


 + =
∂z 2 ∂x 2 ∂x∂z
Other three equations can be obtained with a similar procedure:
 2  
 ∂ x = ∂

 −
∂yz
+
∂xz
+
∂xy



 ∂y ∂z ∂x ∂x ∂y ∂z
 ∂2



∂xz ∂xy ∂yz

y
= − + + (20)
 ∂x∂z
 ∂y ∂y ∂z ∂x

 ∂ 2 z
  
 ∂ ∂xy ∂yz ∂xz
= − + +



∂x∂y ∂z ∂z ∂x ∂y
Linear elastic behaviour for isotropic materials

  
In elastic isotropic materials the 
 x = 1/E σx − ν(σy + σz )
strains, up to stresses near the yield 
  
y = 1/E σy − ν(σx + σz )


condition, are linear functions of the 

z = 1/E σz − ν(σx + σy )

state of stress as: 




 2(1 + ν)
γxy = τxy (21)

 E
where E is the Young’s modulus, ν is


 2(1 + ν)
γyz = τyz
E

the Poisson’s ratio and G = 2(1+ν) is




 E
the shear modulus.


 2(1 + ν)
γxz = τxz

The equations are the Hooke’s law. 
E
Linear elastic behaviour for isotropic materials

  
In elastic isotropic materials the 
 x = 1/E σx − ν(σy + σz )
strains, up to stresses near the yield 
  
y = 1/E σy − ν(σx + σz )


condition, are linear functions of the 

z = 1/E σz − ν(σx + σy )

state of stress as: 




 2(1 + ν)
γxy = τxy (21)

 E
where E is the Young’s modulus, ν is


 2(1 + ν)
γyz = τyz
E

the Poisson’s ratio and G = 2(1+ν) is




 E
the shear modulus.


 2(1 + ν)
γxz = τxz

The equations are the Hooke’s law. 
E

In the case the temperature is increased (or decreased) from a reference temperature
To , the first three equations have to be modified in:
 
 = 1/E σx − ν(σy + σz ) + α · ∆T

 x

 
y = 1/E σy − ν(σx + σz ) + α · ∆T (22)

  
z = 1/E σz − ν(σx + σy ) + α · ∆T

where ∆T = T − To and α is the coefficient of thermal expansion.


Constrained deformation
z
σz

Let us consider a material loaded in z


direction, that is constrained such that
its deformations is constrained along y
material
axis and free along x axis (such as a y

material within a rigid die). die

If we impose y = 0 inthe second expression of (21) we obtain: σy = νσz .


The strain along z then becomes:

1 1 − ν2
z = (σz − ν 2 σz ) = σz
E E
The stiffness of the material along z is then:

σz E
E0 = =
z 1 − ν2
So if the deformation/displacement is constrained, then it results in a increased
stiffness when the material is loaded in one of the other directions.
This situation is called plane strain.
Principal directions for strains and stresses

Dealing with strain-gage measurements, one can’t help wondering if the direction of
principal strains is the same as the direction of principal stresses.
If we take a 2D deformation, we can take Eq. (13):
x + y x − y γxy
X 0 = + · cos 2θ + sin 2θ (23)
2 2 2
If we express the strains in terms of stresses through the Hooke’s law for an isotropic
material, it is then possible to calculate the values of θp for which dX 0 /dθ = 0 (we
are then looking for the direction of principal strains). Solving we then obtain:

dθ 2 1+ν
E
τxy 2τxy
=0 → tan 2θp = 1+ν
=
dθ E
(σx − σy ) (σx − σy )

This means that the direction of principal stresses coincides with direction of principal
strains.

This result is valid for any elastic isotropic material.


Hydrostattic stress and volumetric strain

If we introduce Eq..(21) in the definition of volumetric strain V we obtain:

1 − 2ν
V = (σx + σy + σz ) (24)
E
It can be so seen that the volumetric strain is controlled by I1 (that is the hydrostatic
stress σh = I1 /3).
z

C+dC
C

y
A
B A+dA

B+dB
x

It is worth remarking that when ν = 0.5 (in the plastic material behaviour) V → 0.

You might also like