Sie sind auf Seite 1von 27

For uniaxial testing condition

Load P
Engineering stress e  
Initial cross sec tion area A0 Load
change in length l
=P
Engineering strain, e   e  
Initial length l

Load P
True stress,   
Actual cross sec tion area A l
dl l 
True strain, d  ;   ln  i 
l  l0 

l1 and l0 are the instantaneous and initial gage length of the sample.

For polymers stretch ratio is used for expressing the deformation.

Stretch ratio = (1+e)

Relation between true stress and engineering stress,    e 1  e 

Relation between true strain and engineering strain,   ln 1  e 

Shear stress and shear strain

dl τ

Shear Load P
Shear stress,   
Area A

dl
Shear strain, d   tan 
l
σy , τy

E, G
σ, τ

ε, γ
In the elastic regime the stress and strain are linearly related.

For uniaxial stress state,   E , E is called Young’s modulus or elastic modulus

For shear stress state,   G , G is called shear modulus

For an isotropic material, the elastic and shear modulus are related to each other.

E  2G 1    , where ν is the Poisson’s ratio

First principle explanation of Elastic modulus

Interatomic energy between two adjacent atoms is often considered as the sum of an attractive and
a repulsive term.

Repulsive Energy
Interaction energy, Ui

r0
r

Total Energy

Attractive Energy
Force

Interatomic spacing, r

(dF/dr) α (E)
A B
Ui     attractive  repulsive
rm rn

U i
Interatomic force per bond, dF 
r

Assuming the number of atoms per unit area to be N, the stress can be expressed as

dσ = N.dF

dr
If we ascribe an area (r0x r0) per atom, then N=(r0)-2. The strain can be represented as d  .
r0

Using the above equations for stress and strain, the Young’s modulus can then be expressed as

d 1 dF 1 dU i2
E  
d  0 r0 dr r  r0 r0 dr 2 r  r0

Further using the conditions (dUi/dr)=0 at r=r0,

Am(n  m)
E
r0m  3

An  1
For ionic solids the attraction force is columbic in nature and m=1. Thus E 
r04

m
Bulk Modulus, K  , where σm=1/3(σ11+σ22+σ33) is the hydrostatic stress and Δ=(ε11+ε22+ε33) is

the volumetric strain (ΔV/V).

K, G, E and ν are related with each other.

E 9K 31  2 K E 1  2G / 3K 
K ; E ; G ; K ; 
31  2   3K  21     3E  2  2G / 3K 
1   9 
 G  G 
Figure shows the log-log plot of Bulk modulus (K) and atomic spacing (r0). [Ref-Courtney, mechanical
behaviour of Materials]
Figure shows the variation Bulk modulus with melting point and Young’s modulus with atomic
number along a row of elements in the periodic table. [Ref-Courtney, mechanical behaviour of
Materials]
Relation between stress and strain for more complex state of stress

σ33
x3 τ32

τ31

τ12
τ13
τ23
x2
σ11

τ21
σ22
x1

Two assumptions are used for describing the relation between stress and strain of an isotropic solid.

(1) The longitudinal stresses (σ11, σ22 and σ33) causes longitudinal strains (ε11, ε22 and ε33) while
shear stresses (τ12, τ21, τ13, τ31, τ23 and τ32) leads to shear strain (γ12, γ 21, γ 13, γ 31, γ 23 and γ32).
(2) Principle of superposition holds for longitudinal strains, i.e., the total strain in one direction
is the sum of the strains generated by the various stresses along that direction.

For an isotropic solid, the stress σ11 will leads to longitudinal strain ε11 and transverse strains ε22 and
ε33.

 11  11
11  ;  22   33  
E E

Similarly σ22 and σ33 will introduce following strains

 22  22
 22  ; 11   33  
E E

 33  33
 33  ; 11   22  
E E

Adding the longitudinal strains along x1, x2 and x3, we get

11 
1
 11   22   33 
E
 22   22    11   33 
1
E
 33   33    11   22 
1
E
 12  23  13
and  12  ;  23  ;  13 
G G G

Note that for a body in equilibrium state the stress tensor is a symmetric tensor. So σ12=σ21, σ13=σ31
and σ23=σ32.

The relation between stress and strain can also be expressed in a notation form.

 1    
 ij    ij    kk ij
 E  E

 E  E
 ij    ij   
 1   1   1  2  kk ij
In the above equations i,j varies between (1,2 and 3), σkk=(σ11 + σ22 + σ33), εkk=(ε11 + ε22 + ε33) and
δij=1 (for i=j) and 0 (for i≠j).

For Plane strain deformation condition (example rolling) ε3=0.

Thus σ33=ν(σ11+σ22), stress exists even though ε33=0. The other two strain components can be
expressed as

1 
1
E
 
1  2  1  1    2 
 
1  1  2  2  1    1
1
E

For Plane stress deformation condition, σ3=0 (Example Pressurized vessel)

1 
E
1  v 2 

1  v2 
2 
E
 2  v1 

1  v2 
 E 
Further for isotropic thin films on substrate one can assume ε1=ε2 giving  1   2   1 . This
1 v 
expression is called Stoney equation and is often used for estimation of stresses in thin films. The
strain component ε1 is obtained from the curvature of the substrate. If R is the radius of curvature
then

ES hS2
R , s, f stands for substrate and film and h is their respective thickness.
61   f h f

E/(1-ν) is called biaxial modulus.


Elastic strain energy stored in an isotropic solid

1
U 0   ij ij
2
U 0   1111   22 22   33 33   12 12   13 13   23 23 
1
2

U0 
1
2E
 
 x2   y2   z2   x y   y z   x z  
E
1 2
2G

 xy   yz2   xz2 
1
 1
 
U 0  2  G  x2   y2   z2  G  xy2   yz2   xz2
2 2

U 0 U 0
Then    2G x   x and  x
 x  x

E
Note   is called Lame’s constant.
1   1  2 
Description of strain at a point
C’

dx D D’
x C
A B
exy
u u+(∂u/∂x)dx

P eyx
B’
A’ B’
dx + (∂u/∂x)dx
A B
One dimension strain Angular distortion of an element

Figure taken from Dieter, Mechanical Metallurgy.

For one dimension ex= (∂u/∂x) and u = exx.

For a general three dimensional element the displacement vector with components (u,v,w) can be
u  exx x  exy y  exz z
written as v  eyx x  eyy y  eyz z or ui = eijxi
w  ezx x  ezy y  ezz z

The shear displacements (exy=DD’/DA=∂u/∂y; eyx=BB’/AB=∂v/∂x) are positive if they rotate a line
from one positive axis towards another positive axis.

The relative displacement tensor can be written in matrix form as


 u u u 
 
exx exy exz   x y z 
   v v v 
eij  eyx eyy eyz    
e e e   x y z 
 zx zy zz   w w w 
 
 x y z 

The relative displacement tensor contains contributions from strain tensor and rotation tensor.

eij 
1
eij  e ji   1 eij  e ji    ij  ij
2 2

 u 1  u v  1  u w 
      
 x 2  y x  2  z x 
 xx  xy  xz 
   1  v u  v 1  v w  
 ij   yx  yy  yz         
     2  x y  y 2  z y  
 zx zy zz   
 1  w  u  1  w  v  w 
 2  x z  2  y z  z 
 

 1  u v  1  u w 
0      
2  y x  2   z  x 
0  xy  xz  
   1  v u  1  v w  
ij   yx 0  yz       0    
  0   2  x  y  2  z y 
 zx zy   
 1  w  u  1  w  v  0 
 2  x z  2  y z  

 ui u j 
Note that the engineering shear strain,  ij      2 ij is not a tensor quantity.
 x x 
 j i 

The strain tensor can be further separated in two parts.

   
 ij   ij   m    ij   ij    ij
 3  3

ε’ij is called the deviatoric strain and is involved in shape change, while εm is called the hydrostatic
strain is responsible for change in volume. Here Δ=ε11 +ε22 + ε33.
Graphical solution of a Biaxial stress state: Mohr’s circle

For two dimensional stress state if the normal and shear stress components are described by (σx , σy)
and τxy with reference to orthogonal coordinate axis x-y, then for another orthogonal coordinate axis
(x’, y’) rotated by angle θ, the normal and shear stress components can be represented as

n 
1
 x   y   1  x   y cos 2   xy sin 2
2 2
 n    x   y sin 2   xy cos 2
1
2

The above equations can be rearranged and combined in a form similar to the equation of a circle.

2 2

 n 
1
 x   y 
   n
2

1
  x   y    xy 2
 2  2 

This representation of normal and shear stress on any oblique plane in two dimension is called Mohr
1/ 2
 1 2 
circle. The centre of the circle is at   x   y , 0 and its radius is  2  x   y    xy 
1  2
.
2   
Note that for a small material element at equilibrium, τxy=τyx.

For a rotation of θ degrees of coordinate system in physical space, the stress states are rotated by 2θ
in Mohr circle space. The sense of rotation remains the same for both the space. The orthogonal
planes for which τ=0, are called as the principal planes and the corresponding normal stresses on
them are called the principal stresses σ1 and σ2 (by convention σ1 > σ2). The planes with maximum
shear stress are situated at 45ᵒ to the principal directions.

 max   1 x  y   x   y 
2 1/ 2

 
 
   2

 min   2 2 

2 
xy


and

1/ 2
  x   y 2 
 max       xy2 
 2  

State of stress in three dimensions

For any oblique plane in three dimensional space, with its normal vector having direction cosines (l,
m, n) with respect to the orthogonal coordinate system (x, y, z), a non-trivial solution to the normal
   x    yx   zx
stress (σ) exist if   xy     y   zy  0 , where (σx, σy, σz) and (τxy, τyz, τzx, …) are the normal
  xz   yz    z 
and shear stresses, respectively.

The above determinant can be expanded as

 3   x   y   z  2   x y   y z   z x   xy2   yz2   zx2  


  
x y z 
 2 xy yz zx   x yz2   y zx2   z xy2  0

The first term in parentheses (σx+ σy+ σz) is called the first invariant. The coefficient of σ term is
called the second invariant while the last term in the parantheses is the third invariant of the stress
tensor. For any change in the orientatoin of the coordinate axes these inavriants do not change.
Thus, with regard to the first invariant the sum of normal stress components for any orientation of
the coordinate system remains the same.

For a general case, the oblique plane will have both a normal (σ) and a shear (τ) stress component.
The net stress (S) on the oblique plane is related with (σ, τ), its components along the orthogonal
coordinate axes (Sx, Sy, Sz) and the general stress tensor (σx, σy, σz, τxy, τyz, τzx, …) by following
equations.

S2   2  2

S 2  S x2  S y2  S z2
S x   xl   yx m   zx n
S y   xyl   y m   zy n
S z   xzl   yz m   z n

  S xl  S y m  S z n

   xl 2   y m2   z n2  2 xylm  2 yz mn  2 zx nl

For a plane with respect to the principal axes

 2  1   2 2 l 2m2   2   3 2 m2n2  1   2 2 l 2n2

The principal shear stresses occur for the planes bisecting the principal directions.

Tensor notation for stress and strain

Stress and strain are second rank symmetric tensors, i.e., the a12=a21, a13=a31 and a23=a32 where aij
represents the component along ith row and jth column.

For a transformation of coordinate axes from (x, y, z) to (x’, y’, z’) the components of a vector (first
rank tensor) can be represented in notation form as T’I=AijTj. Here i is the free suffix and j is the
dummy suffix and indicates summation over j=1,2,3. Thus, in the expanded form T’ can be
represented as

T '1  A11T1  A12T2  A13T3


T '2  A21T1  A22T2  A23T3
T '3  A31T1  A32T2  A33T3

 A11 A12 A13 


 
The matrix A  A21 A22 A23 is called the transformation matrix and the individual elements aij
 
 A31 A32 A33 
represent the direction cosines between the ith element of new coordinate system and jth element of
the old coordinate system.

For a second rank tensor like stress and strain, the transformation to new coordinate system is
represented as

 '  AAT and  '  AAT

In the notation form this can be expressed as σ’ij = aikajlσkl and ε’ij=aikajlεkl .

The concept of transformation matrix is often used to define the crystallographic texture of
materials. For example figure below shows the orientation relation between the sample coordinate
system (consisting of rolling direction (RD), transverse direction (TD) and normal direction (ND)) and
crystal coordinate system of a grain (the three cube axis [100], [010] and [001]). The transformation
matrix between these coordinate system can be written as
CC  g . CS
cos 1 cos 1 cos  1 
g  cos  2 cos  2 cos  2 
cos  3 cos 3 cos  3 

Here CC and CS are vectors represented in terms of crystal and sample coordinate system.

Figure. Relationship between sample coordinate axes and crystal coordinate axes in terms of angle
between them..

Principal strains

Similar to stress, it is possible to define a system of coordinate axes along which there are no shear
strains. These axes are the principal strain axes. For an isotropic body the direction of principal
strains coincide with principal stress directions.

The three principal strains are the roots of the cubic equation  3  I1 2  I 2  I 3  0 , where

I1   x   y   z
1 2
I 2   x y   y  z   x z 
4
 
 xy   yz2   zx2
1
4
1

I 3   x y z   xy yz zx   x yz2   y zx2   z xy2
4

The directions of the principal strains are obtained from the three equations

2l    x   m yx  n zx  0
l xy  2m   y   n zy  0
l xz  m zy  2n   z   0

And the principal shearing strains are γ1=(ε2-ε3), γ2=(ε1-ε3) and γ3=(ε1-ε2).
Hydrostatic and Deviator component of stress and strain

   m   'ij

The hydrostatic or mean stress is given as  m 


 kk

 x  y z

 1   2   3 
3 3 3

1
The stress tensor can be written as  ij   'ij   ij kk
3

2 x   y   z
 xy  xz
3
2 y   x   z
 'ij   yx  yz
3
2 z   x   y
 zx  zy
3

Since σ’ij is a second rank tensor, it has principal axes. The principal values of the stress deviator are
the roots of the cubic equation σ’3-J1σ’2-J2σ’-J3=0. J1, J2 and J3 are the invariants of deviator stress
tensor.


Similar to stress  ij   'ij   ij where Δ=(εx+εy+εz) is the volumetric strain.
3

 x   m  xy  xz

 'ij   yx  y   m  yz m 
3
 zx  zy  z   m
and

Anisotropy of Elastic behavior

The elastic stress (σ) and strain (ε) tensor are related to each other by fourth rank stiffness (C) and
compliance (S) tensor.

 ij  Cijkl kl
 ij  Sijkl kl

Thus for an anisotropic elastic solid both normal strains and shear strains are capable of contributing
to a normal stress.

Due to the symmetric nature of stress and strain tensor the 81 components of Cijkl and Sijkl reduces to
only 36 independent components. Moreover, as strain and stress are work conjugates, the 36
components further reduces to only 21 independent components for a general crystal structure
without any symmetry.
For example: σij = σji → Cijklεkl=Cjilkεlk (since εkl=εlk)→Cijkl=Cjilk

Similarly One can prove that Cijkl=Cjikl , Sijkl=Sjilk and Sijkl=Sjikl.

In relation to stress and strain being work conjugates,

U
  1  C111  C12 2  C13 3  C14 23  C15 13  C16 12
1
 2U
 C12
 21

 2U
Similarly  C21
1 2

 2U  2U
Then assuming equivalence of mixed partials,  so C21=C12.
1 2  21

By using the convention  11   1 ,  22   2 ,  33   3 ,  23   4 ,  13   5 ,  12   6 and


11  1 ,  22   2 ,  33   3 ,  23   23 / 2   4 / 2, 13   13 / 2   5 / 2, 12   12 / 2   6 / 2

the stress and strain relation can be written as

 1  C11 C12 C13 C14 C15 C16  1 


  C C C C C C   
 2   21 22 23 24 25 26   2 
 3  C31 C32 C33 C34 C35 C36   3 
   
 4  C41 C42 C43 C44 C45 C46   4 
  C C C C C C   
 5   51 52 53 54 55 56   5 
 6  C61 C62 C63 C64 C65 C66   6 

For cubic system the stiffness matrix can be written as

C11 C12 C12 0 0 0 


. C C 0 0 0 
 11 12 
. . C11 0 0 0 
 
. . . C44 0 0 
. . . . C44 0 
 
. . . . . C44 

Note that the matric is a symmetric one. For isotropic materials (most polycrystalline aggregates can
C11  C12
be treated as such): C44  and thus only 2 independent components C11 and C12 are
2
C  C12
required. For anisotropic systems, C44  11 , and an anisotropy ratio (called the Zener
2
anisotropy ratio) a is defined.
2C44
A
C11  C12

Anisotropic ratio for some metals.

Material W Al Ta Fe Ni Ag Cu
A 1.0 1.22 1.56 2.43 2.51 3.01 3.21

In a cubic material, the elastic moduli can be determined along any orientation, from the elastic
constants, by the application of the following equation:

1  1 

 S11  2 S11  S12  S44  li21l 2j 2  l 2j 2lk23  li21lk23 
Eijk  2 

where Eijk is the Young’s modulus in the [ijk] direction; li1, lj2 and lk3 are the direction cosines of the
direction [ijk]. The Eijk can be also represented in terms of E<100> and E<111>, the modulus along <100>
and <111> crystal directions.

 1 1  22
1

1
 3  
 li1l j 2  l 2j 2lk23  li21lk23 
Eijk E100   E100  E111  

For face-centred-cubic and body-centred-cubic metals, E111>E110>E100, while for simple cubic
material, like cubic zirconia, E100>E110>E111.
Figure. Dependence on orientation of Young’s modulus for monocrystalline (a) Copper, (b) cubic
zirconia and (c) zirconium. (From Meyers and Chawla).

Elastic properties of Polycrystals

(1) Iso-strain model (Voigt Average)

Deformation of all the grains or phases (in composites) is equal to the applied deformation. This
assumption satisfies the strain compatibility condition between all the grains (or phases) of a
material, but need not satisfy the stress equilibrium condition. It gives an upper bound to the elastic
modulus.

If Ei and Vi are the elastic modulus and volume fraction of different grains or phases (in composites)
along a direction, the net elastic modulus of the polycrystalline materials or composite can be
expressed as

E   EiVi

Similarly for shear modulus G  GV


i i

(2) Iso-stress model (Reuss Average)

The stress state for all the grains or phases (in composites) is equal to the applied stress state. This
assumption satisfies the stress equilibrium condition between all the grains (or phases) of a material,
but need not satisfy the strain compatibility condition. It gives a lower bound to the elastic modulus.

1 V
  i , the terms have similar meaning as explained earlier.
E Ei

In ceramics, the elastic moduli are strongly influenced by the presence of porosity. If one assumes
the law of mixtures for the porosity, then, as a first approximation, one has

E  E A (1  f B )  EB f B

where f is the volume fraction of a phase and the subscripts A and b denote the two phases.
However, if phase B is the pore and denoting the pore fraction by p, one has

E  E0 (1  p )

The change in Young’s modulus with porosity has been empirically expressed by Wachtman and
Mackenzie [Mechanical Properties of ceramics, ed J.B. Wachtman, NBS Special Publication 303
(1963) p.139; J.K. MacKenzie, Proc. Phys. Soc., B63 (1950) 2],


E  E0 1  f1 p  f 2 p 2 
where p is the porosity and f1 and f2 are constants. For spherical voids, Mackenzie found that f1 and
f2bare equal to 1.9 and 0.9, respectively, for a Poisson’s ratio of 0.3.
The presence of microcracks in the ceramics decreases the stored elastic energy and reduces the
effective Young’s modulus.

Viscoelasticity/ Anelasticity

Glasses or amorphous materials like polymers show the phenomenon of time dependent strain,
called viscoelasticity or anelasticity. Polymers, polymer solutions and dispersions, metals at very high
temperatures, and amorphous materials (like Bulk metallic glasses, BMG) show viscoelastic
behaviour – that is, characteristics intermediate between perfectly elastic (σ=Eε) and perfectly
viscous (σ=η(dε/dt)n, where η is the viscosity) behaviour.

A viscoelastic substance has a viscous and an elastic component. Figure below shows the stress-
strain curve of an ideal elastic material. The load and unload curves are the same, and the energy
lost as heat per cycle is zero. In practice, there is always an anelastic (i.e, a time-dependent)
component present in practice, with the result that the unload curve does not in fact follow the load
curve. Thus, the energy equal to the shaded area is dissipated in each cycle.

In order to characterize the viscoelastic behaviour of a material, the material is sinusoidally


deformed, and the resulting stress is recorded. For an ideal elastic material, the stress and strain are
in phase, and the phase shift δ=0. For an ideal viscous material, the stress and strain are 90ᵒ out of
phase (i.e., δ = 90ᵒ). For a viscoelastic material a combination of an ideal elastic response and an
ideal viscous response is present. The stress and strain response of a viscoelastic material can be
represented as

   0 sin t
   0 sin t   

In the equation for stress, δ is the phase angle or phase lag between the stress and strain. For such
behaviour two moduli are defined
 
E   0  cos 
 0 
 
E   0  sin 
 0 

where E’ is the tensile storage modulus and E’’ is the tensile loss modulus.

The loss tangent is defined as

Energy loss E


tan   
Energy stored E 

Flow, Yield and Failure Criteria (Empirical relationships)

The meaning of failure may change under different circumstances. In general the failure of a
material can be classified into three groups.

Yielding – For well annealed metals percolation of plastic deformation across the gage section
signifies yielding

Flow – In previously deformed materials onset of plastic flow is marked with percolation of plastic
deformation across the gage section

Failure/Fracture – Applicable to brittle materials, in which the limit of elastic deformation coincide
with failure

Two concepts must be satisfied for any general failure criteria

(1) Uniaxial experiments are easy to perform but the main focus of a failure criterion is to predict the
onset of plastic deformation or fracture in a complex state of stress when one knows the flow stress
of the material under uniaxial tension.

(2) The hydrostatic stress does not contribute to plastic deformation of metallic materials, under
circumstances where dislocation plasticity dominates other deformation processes.

Three failure criteria are listed below.

(1) Maximum-stress criterion (Rankine)

Plastic flow takes place when the greatest principle stress in a complex state of stress reaches the
flow stress in uniaxial tension. Since σ1 > σ2 > σ3, we have

σ0 (tension) < σ1 where σ0 is the flow stress of material.

The greatest weakness of the criterion is that it predicts plastic flow of a material under a hydrostatic
state of stress. Thus this criterion does not hold for metallic systems deforming with dislocation
plasticity.

(2) Maximun shear stress criterion (Tresca Criterion, 1864,1867)


Plastic flow starts when the maximum shear stress in a complex state of deformation reaches a
value equal to the maximum shear stress at the onset of flow in uniaxial tension (or compression).

1   3
 max 
2

For uniaxial test, σ1=σ0, σ2=σ3=0 so, τmax=(σ0/2).

Thus, the Tresca yield criterion can be re-stated as  1   3   0 .

Under hydrostatic condition, σ1=σ3 so there so no yield.

(3) Maximum distortion energy criterion (von Misess 1913)

Plastic yielding in a materials occurs when the second invariant of the stress deviator J2 exceeds
some critical value.

J2  k 2

where J 2 
1
6

1   2 2   2   3 2   3  1 2 . 
For yielding under uniaxial tension test, σ1=σ0, σ2=σ3=0, so we have J2=(1/3)σ02 or k = (1/√3)σ0
=0.577σ0. Using this critical value of J2 the von Mises yield criterion becomes

0 
1
2
( 1   2 ) 2   2   3   ( 3   1 ) 2
2

1/ 2

Or for a general state of stress

0 
1
2
 
 x   y 2   y   z 2   z   x 2  6  xy2   yz2   xz2  1/ 2

The criterion was stated by von Mises without a physical interpretation. It is now accepted that it
expresses the critical value of the distortion (or shear) component of the deformation energy of a
body.

Derivation of distortion energy (Hencky 1924)

Expressing the total strain energy in terms of the principal stress components

U0 
1
2E
 
 12   22   32  2  1 2   2 3   3 1  
Expressing the above equation in terms of the invariants of stress tensor

U0 
1 2
2E

I1  2 I 2 1    
Further expressing the total energy in terms of the bulk modulus, K, and the shear modulus, G by using the
relations
9K 1  2G / 3K 
E 
 3K  2  2G / 3K 
1  
 G 

U0 
I12

18 K 6G
1 2
I1  3I 2 
Thus the total elastic strain energy can be split into a term depending on change of volume and a term
depending on distortion.

U 0 distortion  1
6G

 12   22   32   1 2   2 3   1 3 

U 0 distortion  1
12G

( 1   2 )2   2   3   ( 3   1 )2
2

For a uniaxial state of stress, σ1=σ0, σ2=σ3=0

U 0 distortion  1
2 02
12G
Substituting this critical value of distortion energy for yielding the yield criterion becomes

0 
1
2
( 1   2 ) 2   2   3   ( 3   1 ) 2
2

1/ 2

which is similar to criterion derived with critical J2 approach.

For a state of pure shear, σ1=(-σ3), σ2=0. So according to the von Mises criterion, k=σ1. So the yield
stress σ1 under pure shear condition will be 0.577 times the yield stress under uniaxial tension
condition (σ0).

Some important points regarding the von Mises criterion:

(1) Since it involves (σ1-σ2), (σ2-σ3) and (σ3-σ1), it is independent of hydrostatic stress and is a
function of the three principal shear stresses.

(2) It is independent of the sign of the stresses as square of difference between the principal stresses
are involved.
Octahedral shear stress and shear strain

An octahedral plane makes equal angle with the principal stress directions.

The normal to the octahedral planes make an angle 54ᵒ44’ with the three principal stress axis.

The stress acting on each octahedral plane can be resolved into normal (σoct) and shear (τoct)
components.

1   2   3
 oct   m
3

Thus the normal component of stress on the octahedral plane is equal to the hydrostatic stress.
Thus, it cannot produce plastic yielding.

 oct 
1
3

1   2 2   2   3 2   3  1 2  1/ 2

For an uniaxial stress, σ1=σ0, σ2=σ3=0

2
 oct  0
3

Thus if we represent the yield criterion in terms of the critical shear stress on the octahedral plane,
we again get the von Mises criterion as

0 
1
2
( 1   2 ) 2   2   3   ( 3   1 ) 2
2

1/ 2

The normal and shear components of strain on the octahedral plane are given as

1   2   3
 oct 
3

 oct 
2
3

1   2 2   2   3 2   3  1 2  1/ 2

Where ε1, ε2 and ε3 are the principal strains.


Graphical representation of Yield criteria

For a plane stress condition, σ3=0, the yield criteria can be shown in two dimensional stress space

as below.

The Tresca maximum stress shear stress criterion is more conservative than the von Mises yield
criterion. For any stress state represented by a point inside the (blue) hexagon, no yielding is
expected for either criterion. When the stress state comes on the circumference of the hexagon, the
Tresca criterion is satisfied and the materials should start to yield according to Tresca criterion.
However, if the stress state is still inside the elliptical (orange) region, the von Mises yield criterion is
not satisfied. Only once the stress state represents a point on the circumference of the ellipse the
material will start to yield according to von Mises criterion. For all the points outside the shaded
region the material will show yielding.

Note that both Tresca and von Mises criterion gives same yield stress for uniaxial tension, uniaxial
compression and balanced biaxial stress (σ1=σ2). The greatest divergence between the two criterion
occurs for pure shear (σ1= -σ2) with σvon Mises≃ 1.15 σTresca. In reality the yield surface can be
anisotropic due to presence of crystallographic texture in the material or other direction properties
(as in composites).

For a plane stress condition, the curve along the circumference of the hexagon and the ellipse are
called the yield locus for Tresca and von Mises criterion. The yield locus for a plane stress condition
can be determined by loading a hollow tube with axial and torsion stress. Alternatively, a hydrostatic
pressure may be introduced to produce a circumferential hoop stress in the tube [Refer to Dieter for
details].

Generally after yielding the material undergoes strain hardening and thus the flow stress keeps
increasing with applied strain. This is reflected as an increase in the size of the yield locus. The
hardening in materials can be broadly classified as isotropic hardening and kinematic hardening.
For a general state of stress, the yield criterion can be represented geometrically by a cylinder
oriented at equal angle to the principle stress directions. Since the plastic deformation in metals is
generally not influenced by hydrostatic stress, the generator of the yield surface is a straight line
parallel to the hydrostatic stress line, so that the radius of the cylinder is constant for von Mises
criterion. However, if the yield criterion shows a dependence on hydrostatic stress (for eg. in Bulk
metallic glasses, polymers, soil etc.) then the generator of the yield locus is inclined to the
hydrostatic stress line and the yield locus appears like a cone (eg., Drucker-Prager Yield criterion). As
deformation continue the yield surface expands outwards in accordance with characteristics of
hardening (isotropic or kinematic).

Invariants of stress and strain

If the plastic stress-strain curve (the flow curve) is plotted in terms of invariants of stress and strain
then approximately the same curve will be obtained regardless of the state of stress.

Nadai had shown that the octahedral shear stress and shear strain are inavariant functions which
describe the flow curve independent of the type of test. However, the most frequently used
invariant function to describe the plastic deformation is effective stress and effective strain.

 
2
2
 
 1   2 2   2   3 2   3   1 2 1 / 2

d 
3
2

d1  d 2 2  d 2  d 3 2  d 3  d1 2  1/ 2

The effective strain can be further simplified as


2
 
d   d12  d 22  d 32  1 / 2
3 

Or in terms of the total plastic strain as

 p    12   22   32 
2  1/ 2

3 

For tensile test:

1   ; 2   3  0
1  0 ;  2   3  1
1   2   3  0  1  2 2  0
d1  2d 2  2d 3
   ; d  d  1

For pure shear:

 1   2   ;  3  0
  3

Anisotropy in flow behaviour of deformed materials

Often mechanical processing of materials, such as rolling, introduces crystallographic texture which
influences their flow behaviour when tested along different direction. Figure below shows earing
phenomenon in aluminium foils during deep drawing and cracking in outer surface during of bending
of rolled sheets with aligned oxide particles.
Due to the development of specific crystallographic texture in rolled sheets the tensile properties of
samples along different directions may show different flow behaviour (Figure below). In rolled
sheets Lankford parameter provides a measure of this anisotropy.

Lankford coefficient/ value for tensile sample is expressed as

 width
R 
 thickness

where θ is the orientation of the tensile axis with respect to the rolling direction (RD) and εwidth and
εthickness are the plastic strains along the width and thickness of the sample.

Two parameters Ravg and Rp are defined to give an indication of formability and in plane anisotropy in
flow behaviour of rolled sheets.

R0  2 R45  R90
Ravg 
4
R  2 R45  R90
RP  0
2

Note that due to mirror symmetry across the RD R45=R135.

Another way to represent the forming capability of sheet metals is by using the forming limit
diagram (FLD). In this method the sheet is clamped between two dies and a hemispherical punch is
pressed against it. Before deformation a pattern of circles is marked on the surface of the sheet. The
local strains in two orthogonal directions at any point on the sheet surface are estimated from the
change in shape of the circular pattern. Thus, the critical values of orthogonal strains for failure
(either necking or formation of cracks) is determined for various parameters of pressing (eg., friction
coefficient between the sheet metal and the punch) to draw FLD. An example of FLD is shown in
figure below.

Das könnte Ihnen auch gefallen