Nonlinear Solid Mechanics Reader
Nonlinear Solid Mechanics Reader
April 2020
course: 201400042
Contents
Preface vii
1 Introduction 1
1.1 Origins of nonlinear elasticity and inelasticity . . . . . . . . . . . . . . . . . . 1
1.2 The stress–strain curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Metal plasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.2 Time dependent behavior . . . . . . . . . . . . . . . . . . . . . . . . . 3
iii
iv CONTENTS
6 Plasticity theory 51
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.2 One-dimensional plasticity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.2.1 Governing equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
6.3 Plastic deformation at multi-axial stress states . . . . . . . . . . . . . . . . . 53
6.3.1 The yield surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.3.2 Normality rule and the convexity of the yield surface . . . . . . . . . . 54
6.4 Elastic - ideal plastic constitutive equations . . . . . . . . . . . . . . . . . . . 59
6.5 Multi-axial yield functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
6.5.1 Tresca yield criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
6.5.2 Von Mises yield criterion . . . . . . . . . . . . . . . . . . . . . . . . . 62
6.5.3 Comparison of the Tresca and Von Mises yield surface . . . . . . . . . 64
6.5.4 The Hill yield criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
6.5.5 Non-associated flow behavior . . . . . . . . . . . . . . . . . . . . . . . 65
6.5.6 Constitutive equations for elastic-ideal plastic material, using the Von
Mises yield criterion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
6.6 Constitutive equations for a hardening material . . . . . . . . . . . . . . . . . 68
6.7 Hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
6.7.2 Isotropic hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.7.3 Kinematic hardening . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6.8 Elastic-viscoplastic material . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.8.2 Creep in metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.8.3 Visco-plastic material behavior . . . . . . . . . . . . . . . . . . . . . . 80
CONTENTS v
In the course Nonlinear Solid Mechanics, we consider two phenomena that lead to nonlin-
ear relations between forces and displacements: large displacements and nonlinear material
behavior. Both play an important role in material forming processes, in which products are
shaped by plastic deformation. The simulation of forming processes is an important applica-
tion area of the models that are presented in this course. It is the primary research area of the
chair Nonlinear Solid Mechanics. A second application where both large deformations and
nonlinear material behavior are combined is the analysis of rubber or in general of elastomeric
products. For this purpose, hyperelastic models are used, as explained in Chapter 5.
A deeper understanding of the acting process forces is necessary in order to determine the
strength of the tooling and the required press forces. Insight in the material flow is important
in order to assess whether the product will acquire the prescribed shape.
Shape deviations are caused by fracture or tearing during the manufacturing process, but
also by deformations after the forming stage as a result of residual stresses – so-called elastic
spring back.
The methods that will be used for the analysis of forming processes are based on continuum
mechanics. Microscopic material behavior and micro structures are therefore not considered
in this course.
For the analysis of deformation processes, the mathematical description of the material behav-
ior during plastic deformation, by means of constitutive equations, is very important. These
equations present a relation between stresses and deformations. The most general constitutive
equations for anisotropic elastic-plastic material are quite complex and problems with real
deformation processes can often not be solved analytically. Even if we restrict ourselves to
isotropic elastic-plastic behavior analytical solutions are hard to get. Application of numerical
solution methods is then the only feasible option.
With the resulting equations describing the process, it is possible to derive approximating
analytical solutions. With this analytical approximation, a relatively accurate estimation of
the process forces can be made. Analysis of the material flow is not very accurate. A more
precise prediction can be obtained with numerical methods. Therefore, some attention will
be given to the application of the finite element method in the analysis of forming processes.
The relevant quantities in our analysis, like stress and deformation, are mathematically de-
scribed as tensors. In the first chapters, a short introduction to some topics of tensor algebra
is given and the index notation is introduced. Thereafter, we consider the stress tensor and
rate-of-deformation tensor. Next, the constitutive equations for hyperelastic and rigid-plastic
material, the equilibrium equations and the virtual work equation are elaborated. In Chap-
ter 7, we will treat the application of the finite element method. It is followed by a chapter on
the description of damage in a material and a chapter on the stability of plastic deformation
in sheet metal forming.
vii
viii Preface
The course, and also this reader, builds upon previous courses Plasticity and Creep and Solid
Mechanics 2, by Prof. Han Huétink and Dr. Timo Meinders.
Chapter 1
Introduction
1
2 Introduction
voids in ductile materials is referred to as the damage mechanism. After coalescence usually
microcracks form inside the material which lead to fracture.
s s
σuts
σfrac
σy
σp
(a) e (b) e
Figure 1.1: Typical engineering stress–strain curve for (a) metals (b) rubber-like materials.
value of the (re-)yield stress is lower than the value for the original loading direction. This is
referred to as the Bauschinger effect and it is depicted in Figure 1.2.
e s
(a) (b)
t t
Figure 1.3: Observed (a) creep at constant stress and (b) relaxation and recovery at constant
strain.
Chapter 2
In this chapter tensor notation is introduced that generalizes the concept of vectors. Although
this notation may seem cryptic at first sight, it is very strong in describing complex relations
in a concise manner. We will need this notation in the following chapters to avoid lengthy
equations for general 3-dimensional stress-states and deformations.
✯
a
First of all, consider the vector a as drawn in Figure 2.1. We will assume that this vector
represents a physical quantity, like e.g. a displacement, a velocity or a force. The physical
meaning is independent of our (arbitrary) choice of a coordinate system and it is represented
by the arrow.
Two different notation schemes are used in this reader: index-free tensor notation and index
notation. In index-free tensor notation vectors are denoted by bold letters1 , like a. In index
notation, the coordinates in a specific vector basis are used to denote a vector. This is close
to how we expressed a vector in linear algebra, using a column of coordinates:
a1
{a} = a2 (2.1)
a
3
1
In handwriting normal and bold symbols are hard to disinguish and we will put a tilde underneath the
symbol, like a.
˜
5
6 Vectors and tensors
We will restrict ourselves to Cartesian coordinates in R3 (Euclidean space), with unit base
vectors e1 , e2 , e3 as represented in Figure 2.2. For an arbitrary vector a with components
(a1 , a2 , a3 ):
a = a1 e 1 + a2 e 2 + a3 e 3 (2.2)
or
3
X
a= ai e i (2.3)
i=1
We see that in (2.3) the summation takes place over the index that appears twice after the
summation sign for the coordinate directions 1, 2 and 3. In the sequel we will omit the
summation sign and its bounds, and implicitly take the sum over the indices that appear
twice:
a = ai e i dummy-index (2.4)
This is called the summation convention or Einstein convention. For the double index, it is
irrelevant whether it is i, j, k or any other symbol and it is therefore also called a dummy-
index. If a double index occurs it means that the value of the index is subsequently substituted
by 1, 2 and 3 and that the sum of the results of these three substitutions is taken.
In index notation, the vector is commonly denoted just by the components ai , where it is
implicitly assumed that i ∈ {1, 2, 3} and that these are the components with respect to a
Cartesian basis {e1 , e2 , e3 }. So formally, a is a vector and ai is a set of components of a
vector and they are not equal. The relation between them is given by the three equivalent
equations (2.2), (2.3) or (2.4). In practice, we often talk about the vector a in the index-free
notation and the same vector ai in index notation.
When e is a unit vector in the direction of a then
a = kake (2.5)
Vector operations
Scalar multiplication of vector a by α results in a vector b in the same direction but with a
length of αkak:
b = αa or bi = αai (scalar multiplication) (2.6)
2.1 Vectors and vector operations 7
✒
b
✿
a
φ
The sum of two vectors a and b gives a vector c that can be constructed by letting vector b
start at the end of vector a, c is then connecting the start of vector a to the end of b. Vector
addition is commutative.
The inner product or dot product or scalar product of two vectors is a scalar quantity that is
defined as
a · b = kakkbk cos(φ) (inner product of two vectors) (2.8)
where φ is the smallest angle between a and b as presented in Figure 2.3. From this definition,
it is immediately clear that the inner product of two vectors is commutative:
a·b=b·a (2.9)
and that the inner product of two orthogonal vectors is zero. The definition also implies that
the inner product of a vector with itself is the square of the length of that vector:
a · a = kak2 (2.10)
In linear algebra, a distinction was made between row and column vectors. In tensor notation,
a vector has a particular geometrical interpretation, independent of the coordinate system.
The inner product is uniquely defined for two vectors and therefore there is no need to relate
to a representation in a column or row of coordinates or to use the transpose for the first
vector.
The base vectors ei are mutually orthogonal unit vectors and hence, by using definition (2.8)
and (2.10): (
1 if i = j
ei · ej = δij = (2.11)
0 if i 6= j
where we implicitly defined the Kronecker delta symbol δij .
Using the Kronecker delta symbol and the summation convention, we can write the vector
component ai as:
ai = δij aj (2.12)
where j is a dummy index taking the values 1, 2 and 3, but only if i = j, the value of aj
contributes to the result.
The inner product of two vectors a and b can then be written as:
The components of a vector in a particular Cartesian basis can easily be obtained using the
inner product.
ai = a · ei (components of vector a) (2.14)
This follows directly from (2.2) and orthogonality of the base vectors.
The cross product or vector product of two vectors results in another vector:
where the length of c is equal to the area of the parallelogram spanned by a and b:
and the direction of c is perpendicular to the plane spanned by a and b, in the direction given
by the right-hand rule. Because of this right-hand rule, the cross product in not commutative,
but:
a × b = −b × a (2.17)
The triple product of 3 vectors a, b and c, takes the inner product of a with the cross product
of b and c. Since the cross product represents the area of the parallellogram spanned by b
and c and the inner product subsequently defines the size of a in the direction of the normal
to the parallellogram, the triple product represents the signed volume V of the parallellepiped
spanned by a, b and c.
V = |a · (b × c)| (triple product of 3 vectors) (2.18)
The sign of the triple product is positive or negative depending on the orientation of the
vectors with respect to each other.
2.2 Tensors
A vector is also called a first order tensor and a scalar is sometimes called a zero-th order
tensor. We can also define higher order tensors.
First we define the dyadic product or tensor product of two vectors a and b by the dyad2 ab
implicitly by the property
Here, we used the property that a scalar multiplication (with bj ) can be replaced freely within
the product. However this does not hold for the dyadic product ei ej ; the order of vectors
within a dyadic product must be maintained. Although a dyadic product is often just written
as ai bj in index notation (and this is equal to bj ai since the components are scalar values), it
must be reminded that this represents coordinates for a dyad of the vectors ei and ej in that
particular order.
2
We follow the convention that there is no operator symbol between the two vectors to denote the tensor
product. In another widely used notation set this dyad is denoted by a ⊗ b.
2.2 Tensors 9
A second order tensor A is an operator that maps an arbitrary vector b to another vector c.
It can be expressed as a linear combination of dyads and is represented in this reader by bold
letters3 , e.g. A. A second order tensor can be expressed in index notation with reference to
a Cartesian coordinate system as:
A = Aij ei ej (second order tensor A) (2.23)
and the components can be obtained from:
because the inner product ei ·ej results in a scalar (δij ) and can therefore be placed at another
position in the product.
The second order identity tensor 1 maps any vector onto itself:
c=1·c=c·1 (2.27)
By the definition of a second order tensor as linear combination of dyads, it also holds for
second order tensors:
A=1·A=A·1 (2.28)
3
In handwriting we will put a bar underneath the symbol, like A.
10 Vectors and tensors
c = ci ei = c · ei ei = c · 1 ∀c ⇒ 1 = ei ei (2.29)
The components of the second order identity tensor are given by the Kronecker delta symbol:
1 = δij ei ej = ei ei (2.30)
For 2 dyadic products the double contraction or double dot product can be defined by4 :
and likewise for 2 second order tensors: dyadic product between vector a and b
The colon represents a summation over both indices (double contraction) and the result is
a scalar. Because the dot product of two vectors is independent of the coordinate system,
it also holds that the double contraction of two second order tensors is independent of the
coordinate system.
Next, the trace of a second order tensor can be defined as:
which equals the sum of the diagonal terms of the components of A when written as a matrix.
A double contraction of a second order tensor A with the second order identity tensor also
yields the trace of A:
Analogous to the previous discussion we can also define third and fourth order tensors as
linear combinations of tryads of 3, repectively tetrads of 4 vectors. Third and fourth order
tensors will also be denoted by bold symbols5 . The order of the tensor must be derived from
the context. A third order tensor is a linear operator that maps a vector onto a second order
tensor or a second order tensor onto a vector.
A=D·a
(2.35)
Aij = Dijk ak
a=D:A
(2.36)
ai = Dijk Ajk
For a fourth order tensor similar operations hold, e.g.:
A=E:B (2.37)
The fourth order identity tensor is defined by the condition that for any second order tensor
the following relation must hold:
A = 41 : A (2.39)
The fourth order identity tensor can be written as:
4
1 = δik δjl ei ej ek el (2.40)
Note that there is a large difference between the fourth order identity tensor and the dyadic
product of two second order identity tensors. This can easily be shown when contracting
these tensors with an arbitrary second order tensor A:
4
1 : A = δik δjl ei ej ek el : Apq ep eq = Apq ei ej ei ej : ep eq
= Apq δip δjq ei ej = Apq ep eq = A (2.41)
11 : A = δij δkl ei ej ek el : Apq ep eq = Apq ei ei ek ek : ep eq
= Apq δkp δkq ei ei = Akk ei ei = Akk 1 (2.42)
Note that the latter equation yields an identity tensor in which the diagonal terms are mul-
tiplied by value of the trace of the second order tensor A.
Notice that the differentiation ∂/∂xj operates on the vector to the left of the operator, which
cannot easily be seen in the first part of this equation. In tensor notation we also know the
pre-gradient:
∂ ∂aj
∇a = e i aj e j = ei ej = aj,i ei ej (pre-gradient) (2.46)
∂xi ∂xi
12 Vectors and tensors
The pre-gradient is the transposed (conjugated) of the post-gradient. The gradient of a second
order tensor yields a third order tensor:
∂Aij
A∇ = ei ej ek = Aij,k ei ej ek (2.47)
∂xk
The divergence of a vector a yields a scalar (dot product of the gradient operator with a
vector):
∂ ∂
div a = ∇ · a = e i · aj e j = aj ei · ej = ai,i (2.48)
∂xi ∂xi
G = G(t, xi ) (2.50)
The position of the material particle is a function of time: xi = xi (t), therefore G = G(t, xi (t)).
The material derivative—denoted by a superimposed dot—is given by:
dG ∂G ∂G dxi
= Ġ = +
dt ∂t ∂xi dt
∂G ∂G
Ġ = + vi (2.51)
∂t ∂xi
The proof of this equation is given for the case that G is a vector (for simplicity):
f f f
o 1,1 2,1 3,1
→
− n
v · ∇f = vi fj,i = v1 v2 v3 · f1,2 f2,2 f3,2
f1,3 f2,3 f3,3
v1 f1,1 + v2 f1,2 + v3 f1,3
= v 1 f 2,1 + v 2 f 2,2 + v 3 f 2,3
(2.54)
v1 f3,1 + v2 f3,2 + v3 f3,3
f1,1 f1,2 f1,3 v1
←
−
= f2,1 f2,2 f2,3 · v2 = fi,j vj = f ∇ · v
f f f v
3,1 3,2 3,3 3
or in components: Z Z
vi ni dS = vi,i dV (2.56)
S V
or in components: Z Z
Aij nj dS = Aij,j dV (2.58)
S V
Q = Qij ei ej (2.61)
QTij = Q−1 ∗
ij = Qji = ej · ei (2.63)
For vector a:
a = ai ei = a∗k e∗k = a∗k Qik ei
hence
ai = Qik a∗k (2.64)
Multiplication of (2.64) with Q−1
ji gives:
∗ ∗ ∗
Q−1
ji ai = Qij ai = Qji Qik ak = δjk ak = aj
−1
So, for tensor components, the same transformation rule holds as for vector components, but
now applied twice, once for every index.
A reference frame transformation for third order tensors can be written as:
∗
Bijk = QTip QTjq QTkr Bpqr = Bpqr Qpi Qqj Qrk (2.69)
then also:
Pij∗ = A∗ik Bkj
∗
(2.72)
Proof:
∗
A∗ik Bkj = QTil QTkm Alm QTkn QTjp Bnp (2.73)
QTkm QTkn = Qmk Q−1
kn = δmn (2.74)
therefore:
A∗ik Bkj
∗
= QTil QTjp Aln Bnp = QTil QTjp Plp (2.75)
We also have the relation:
Pij∗ = QTil QTjp Plp (2.76)
hence:
Pij∗ = A∗ik Bkj
∗
(2.77)
In general, relations between tensor components remain valid with respect to an arbitrary
base. The transformation rules for tensor components will hardly be used in this course,
because we will use only one coordinate system.
It was already observed below Equation (2.32) that the double dot product of two second
order tensors is does not depend on the coordinate system and therefore also the trace of a
second order tensor is invariant.
The second invariant of a tensor A is defined as:
1
tr2 (A) − tr(A2 )
I2 (A) = 2 (2.79)
where A2 = A · A. Since the dot product of two second order tensors is fully described by
the vectors constituting the dyads and the trace of any second order tensor is invariant, the
result is also invariant. In some books, the second invariant is defined as tr(A2 ) but then the
equations below where I2 is used are not valid.
The third invariant of a tensor A, also called the determinant of A, is defined as:
(A · c1 ) · (A · c2 ) × (A · c3 )
I3 (A) = det(A) = (2.80)
c1 · c2 × c3
where {c1 , c2 , c3 } represents a vector basis. If we consider the definition and interpretation of
the triple product in (2.18) it is clear that the determinant represents the ratio of volume of
the parallellepiped spanned by A · c1 , A · c2 and A · c3 and the volume of the parallellepiped
spanned by c1 , c2 and c3 and this ratio is independent of the coordinate system.
The eigenvalues λ and the eigenvectors n of a tensor A are defined as those scalars and
vectors that satisfy:
A · n = λn with n 6= 0 (2.81)
It can also be written as (A − λ1) · n = 0. A nontrivial solution n 6= 0 is obtained when the
part between brackets is singular, or:
This equation has three roots λ1 , λ2 and λ3 . Because the eigenvalues do not depend on the
coordinate system, they are also invariant.
According to the Cayley–Hamilton theorem a tensor should satisfy its own characteristic
equation, so that (2.83) becomes
A3 − I 1 A2 + I 2 A − I 3 1 = 0 (2.84)
A2 − I1 A + I2 1 − I3 A−1 = 0 (2.85)
Taking the trace of the resulting equation, the following relation between the second and third
invariant can be derived:
I2
tr(A−1 ) = (2.86)
I3
which shows that tr(A−1 ) can also be used as an invariant.
2.8 Tensor derivatives 17
∂I1 (A)
=1 (2.89)
∂A
∂I2 (A)
= I1 1 − AT (2.90)
∂A
∂I3 (A)
= det(A)A−T (2.91)
∂A
Chapter 3
3.1 Kinematics
The kinematics of a deformable body concerns the motion of a material and coordinate system
from an initial state (reference state) to the final situation, see Figure 3.1. In the reference
configuration a material particle is located at position X, the current (spatial) position of the
material particle is defined as x. Several definitions of motion can be used. In the reference
description the only independent variables are the initial position X of a material particle in
an arbitrary reference configuration and time t. In this definition the observer can be thought
of as moving with the material particle. A special case is the Lagrangian description in which
the reference configuration is chosen to be the initial configuration at time t = 0. Another
description of motion is the Eulerian description. Here the only independent variables are the
current location x of a material particle and time t. In this definition the observer is located
at a fixed point in space.
The Lagrangian location vector X determines the initial (reference) position of a particle
within a body at t = 0. In Figure 3.1 the motion history of a particle X is indicated by the
trajectory. The motion of each particle can be described by the Eulerian location vector x
at time t, which is a function of the Lagrangian location vector X (the location in the body)
and the time t:
x = x(X, t) (Langrangian description) (3.1)
19
20 Kinematics and Strain
By means of these coordinates a motion is described by which the particle with the material
coordinates X is transferred to location x at t > 0. The inverse of Equation (3.1) reads:
X = X(x, t) (3.2)
In this Lagrangian or material description the independent variable is the reference position
X. In the Eulerian or spatial description the only independent variable is the fixed position
in space. The material time derivative for a material particle at spatial position x is derived
to be:
The material time derivative is split into two parts. The first part of the right hand side of
Equation (3.4) is the spatial time derivative which is viewed by an observer in a fixed position.
The second term is the convective part; the value of θ at position x changes because another
particle with a different value of θ passes by. The split into a spatial part and a convective
part is necessary if θ is only known in Eulerian coordinates.
In order to describe the deformation of a body, a deformation gradient F is defined. The
deformation gradient F relates the spatial configuration to the reference configuration for a
material particle:
∂x ←
−
F(X, t) = = x∇ 0 (3.6)
∂X t
←−
where ∇0 is the post-gradient with respect to the initial coordinates. The deformation gra-
dient F can be considered as an operator which maps the initial configuration to the current
configuration:
dx = F · dX (3.7)
in which dx and dX represent the same line element in the deformed and undeformed state,
respectively, see Figure 3.2. Generally, F is not symmetric. The displacement u of a particle
is defined as the difference between the initial and the current position (or vice versa) and
reads in Lagrangian coordinates:
The latter can be understood if it is remembered that at one certain position in space, x,
the difference is calculated between the material coordinate of the particle which was initially
at a certain position and which is present at the current time t. Therefore, Equation (3.9)
does not represent the displacement of one specific particle. The gradient of the displacement
reads, in Lagrangian and Eulerian coordinates, respectively:
←
− ∂u ∂x
(Langrangian displacement gradient) u∇0 = = −1 = F−1 (3.10)
∂X ∂X
←− ∂u ∂X
u∇ = =1− = 1 − F−1 (3.11)
∂x ∂x
where 1 is the second order unity tensor.
The determinant of the deformation gradient J represents a change in volume:
dV
J = det F = (3.12)
dV0
in which R is the rotation tensor (which is orthogonal: R−1 = RT ) and U and V the right
and the left stretch tensor, respectively. U as well as V are symmetric and positive definite
tensors. Both decompositions in the latter equation can be considered as a split of F into a
pure deformation part, U or V, and a rigid body rotation R. For the first decomposition in
Equation (3.13) the map consists of a pure deformation, after which a rotation is applied to
the line segment in Figure 3.2. For the second decomposition this sequence is reversed. Both
decompositions are visualized in Figure 3.3.
The right and left stretch tensors are both symmetric and hence have mutually perpendicular
eigenvectors. In the direction of these eigenvectors, the material stretches by a factor, equiva-
lent to the corresponding eigenvalue, respectively λ1 , λ2 and λ3 . In these directions, no shear
deformation is observed. Rotation of the material is represented in R and not in the stretch
tensor. In the undeformed state, the stretch tensors U = V = 1 and the principal stretches
λ1 = λ2 = λ3 = 1.
22 Kinematics and Strain
Proof of *:
(F · dX) · (F · dX) = Fij dXj Fik dXk = Fij Fik dXj dXk = FjiT Fik dXj dXk =
(FT · F) : dX dX (3.17)
The other way around, the square of the initial length can be expressed in terms of the
Eulerian coordinates:
Proof of *:
Assume the following two arbitrary relations: u = A · v and w = B · u. Substitution of the
second equation in the first equation and rearranging gives:
v = (B · A)−1 · w (3.19)
Both equations can also be written as: v = A−1 · u and u = B−1 · w. Again, substitution of
the second equation in the first equation and rearranging gives:
Since Equations (3.19) and (3.20) have to hold for arbitrary w, it is shown that:
Note that the left and the right Cauchy–Green tensors equal the unity tensor in case of no
deformation. In order to overcome this inconvenience, strains are generally defined in terms
of relative changes of the lengths of arbitrary line elements:
and not in terms of relative lengths themselves. For pure rigid body motions the strain
according to this measure then vanishes. Again, this measure can be evaluated in Lagrangian
and Eulerian coordinates. Using the expression for ds2 , Equation (3.16) and recognizing that
ds20 = 1 : dX dX, the deformation measure, Equation (3.22), can be written in Lagrangian
coordinates as:
ds2 − ds20 = (C − 1) : dX dX = 2E : dX dX (3.23)
in which E is the Lagrangian strain tensor, or the Green–Lagrange strain tensor:
E = 21 (C − 1) (3.24)
This tensor vanishes in case of rigid body motions. In terms of the displacements, the Green–
Lagrange strain tensor can be written as (using Equation (3.10)):
h−→ ←
− i
E = 12 (C − 1) = 21 (FT · F − 1) = 12 (∇ 0 u + 1) · (u∇0 + 1) − 1 (3.25)
h ←− →
− →
− ←− i
= 12 u∇0 + ∇0 u + (∇0 u) · (u∇0 )
e = 21 (1 − B−1 ) (3.27)
24 Kinematics and Strain
Like the Green–Lagrange strain tensor, the Euler–Almansi strain tensor vanishes in case of no
deformation and it is geometrically nonlinear. In terms of displacements, the Euler–Almansi
strain tensor can be written as:
h →
− ←
− i
e = 21 (1 − B−1 ) = 12 (1 − F−T · F−1 ) = 12 1 − (1 + ∇u) · (1 + ∇u) (3.28)
h ←− − → →
− ←− i
= 21 u∇ + ∇u − (∇u) · (u∇)
Note that for the Euler–Almansi strain the gradients are defined with respect to the cur-
rent state, contrary to the Green–Lagrange strain. For convenience the strain definitions in
Lagrangian and Eulerian coordinates are summarized in Table 3.1. For small displacements
Lagrange Euler
metric tensor undeformed 1 B−1 = (F · FT )−1
deformed C = FT · F 1
strain tensor E = 12 (C − 1) e = 21 (1 − B−1 )
the geometric nonlinear terms can be neglected. It can also be proven that the gradient
with respect to the initial coordinates almost equals the gradient with respect to the current
coordinates ( ∇0 ≈ ∇ ) in case of small deformations:
∂u ∂u ∂x ∂u ∂u ∂u ∂u ∂u ∂u
= · = · 1+ = + · ≈ (3.29)
∂X ∂x ∂X ∂x ∂X ∂x ∂x ∂X ∂x
This means that for small deformations both the Green–Lagrange and the Euler–Almansi
strain tensors are identical to the classical linear strain tensor ε :
←− →
− ←− − →
ε = 12 u∇0 + ∇0 u ≈ 21 u∇ + ∇u ≈ E ≈ e (3.30)
∂v ∂v ∂X
L= = · = Ḟ · F−1 (3.32)
∂x ∂X ∂x
Like all second order tensors, the velocity gradient can be decomposed into a symmetric and
a skew-symmetric part:
L = D+W (3.33)
3.2 Strain definitions 25
The Green–Lagrange strain tensor and the Euler–Almansi strain tensor are frequently used in
engineering mechanics. A relation between the rate of deformation tensor D and the Euler–
Almansi strain tensor can be found by taking the time-derivatives of Equations (3.22) and
(3.26):
d
( dx · dx − dX · dX) = 2 dv· dx = 2L : dx dx = 2(D+W) : dx dx = 2D : dx dx (3.36)
dt
(Try to proof yourself that W : dx dx = 0 by writing this tensor product in components)
d d
(2e : dx dx) = 2 ( dxi eij dxj )
dt dt
= 2 ( dvi eij dxj + dxi ėij dxj + dxi eij dvj )
= 2 (Lik dxk eij dxj + dxi ėij dxj + dxi eij Ljk dxk )
= 2 (ėij dxi dxj + LTki eij dxk dxj + eij Ljk dxi dxk )
= 2 (ė + LT · e + e · L) : dx dx (3.37)
D = ė + LT · e + e · L (3.38)
This expression is also known as the Cotter–Rivlin derivative of e or the lower convective
time derivative of e. The rate of deformation tensor D relates to the Green–Lagrange strain
tensor via the time derivatives of Equations (3.22) (and thus Equation (3.36)) and (3.23):
Ė = FT · D · F (3.39)
The derivation of this relation is seen easily when writing Equation (3.36) in original coordi-
nates.
Another, very commonly used, strain measure is the logarithmic strain (also known as the true
or natural strain). In case of proportional loading (i.e. principal directions of deformation do
not change), the rate of deformation tensor relates to the logarithmic strain as follows. The
starting point is Equation (3.36):
d d d 2
( dx · dx − dX · dX) = ( dx · dx) = ds = 2D : dx dx ⇐⇒
dt dt dt
2 ds dṡ = 2D : dx dx (3.40)
The right hand side of Equation (3.40) can be written as a function of ds in case of pro-
portional loading only. We consider the deformation of a tensile test. Assume that the first
principal axis is parallel to the tensile direction. Then the right hand side of (3.40) reads:
The logarithmic strain, also called true or natural strain, can be considered as the integral of
the length changes,
l
dl∗ l ∆l
Z
ǫl = = ln = ln 1 + (3.44)
l0 l∗ l0 l0
Finally, Equation (3.42) and (3.43) define the relation between the logarithmic strain and the
rate of deformation tensor for proportional loading:
L = Ḟ · F−1 (3.47)
With Ḟ = Ṙ · U + R · U̇ and F−1 = U−1 · RT the expression for the velocity gradient results
into
L = Ṙ · U · U−1 · RT + R · U̇ · U−1 · RT
or
L = Ṙ · RT + R · U̇ · U−1 · RT (3.48)
but also:
L=W+D
In the case of rigid rotation only: L = Ṙ · RT = W
Even after continued deformation without additional rotation/spin, the rotation tensor may
change whereas the material is not rotating: W = 0 but Ṙ 6= 0 and L = D = Ṙ · RT + R ·
U̇ · U−1 · RT
Conclusion: There exists no simple relation between the multiplicative split of the deformation
tensor F = R · U and the additive split of the velocity gradient L = W + D.
As stated before, there exist no simple relation between the multiplicative polar decomposition
F=R·U
and the additive split
L=W+D
Note: The rotation tensor R may change even in the case of zero spin. In the case of non
proportional deformation; U̇ and U−1 have different principal directions and hence U̇ · U−1
is NOT symmetric.
3.5 Examples
The theory in this chapter will be clarified with three 2-dimensional examples. All examples
are basic deformation modes. The first example is a simple stretching deformation, the
second example will be a simple shear test and the last example comprises a rigid rotation.
All examples make use of a constant strain rate α and a time t.
First, before the examples, the tensors R and U are determined for a 2-dimensional defor-
mation.
Knowing that U is symmetric, the angle β can be determined from the condition U12 = U21
Proof:
If U is symmetric then U12 = U21 so:
∂x ∂x
∂x 1 + αt 0
F= = ∂X ∂Y = (3.64)
∂X ∂y ∂y 0 1
∂X ∂Y
The right and left Cauchy–Green stretch tensors now become:
" #
(1 + αt)2 0
C = FT · F = (3.65)
0 1
" #
(1 + αt)2 0
B = F · FT = (3.66)
0 1
The Green–Lagrange strain tensor relates to the right Cauchy–Green stretch tensor via:
" #
α 2 t2 + 2αt 0
E = 12 (C − 1) = 12 (3.67)
0 0
Note that the same expression can be found when using Equation (3.26).
The Euler–Almansi strain tensor relates to the left Cauchy–Green stretch tensor via:
" #
(α2 t2 + 2αt)(1 + αt)−2 0
e = 21 (1 − B−1 ) = 12 (3.68)
0 0
Note that the same expression can be found when using Equation (3.29).
The classical linear strain tensor reads:
" #
←− →
− αt 0
ε = 21 u∇0 + ∇0 u = 12 ((F − 1) + (FT − 1)) = (3.69)
0 0
Figure 3.5 shows the xx-components of the three strain tensors as a function of the strain
(αt). The velocity field for this problem is:
30 Kinematics and Strain
0.8
0.6
0.2
0
-1 -0.5 0 0.5 1
-0.2
-0.4
E
-0.6
e
-0.8 epsilon
-1
engineering strain
vx = αX (3.70)
vy = 0
∂v ∂v
L = = · F−1 (3.71)
∂x ∂X
" #" # " #
α 0 (1 + αt)−1 0 α(1 + αt)−1 0
= =
0 0 0 1 0 0
" #
1 T
0 0
W= 2 (L −L )= (3.73)
0 0
The time derivatives of the Green–Lagrange strain tensor and the Euler–Almansi strain tensor
are:
Ė = FT · D · F (3.74)
" #" #" #
(1 + αt) 0 α(1 + αt)−1 0 (1 + αt) 0
=
0 1 0 0 0 1
" #
α(1 + αt) 0
=
0 0
3.5 Examples 31
ė = D − e · L − LT · e (3.75)
" # " #" #
α(1 + αt)−1 0 (α2 t2 + 2αt)(1 + αt)−2 0 α(1 + αt)−1 0
= −
0 0 0 0 0 0
" #
α(1 + αt)−3 0
=
0 0
Note that these expressions can also be derived by taking the time derivative of Equations
(3.67) and (3.68).
In this example, the rotation tensor R reads:
" #
1 0
R= (3.76)
0 1
x = X + αtY (3.78)
y = Y
Notice that αt is related to the rotation φ of the vertical material lines in the origin by
αt = tan φ.
The deformation gradient for this deformation pattern is:
" #
∂x 1 αt
F= = (3.79)
∂X 0 1
The Euler–Almansi strain tensor relates to the left Cauchy–Green stretch tensor via:
" # " # " #
1 1 1 0 1 1 −αt 1 0 αt
e = 2 (1 − B ) = 2
−1
−2 =2 (3.83)
0 1 −αt α2 t2 + 1 αt −α2 t2
In the definition of the Green–Lagrange strain and the Euler–Almansi strain tensile or com-
pressive terms evolve in the vertical direction, although only shear deformation is imposed.
The classical linear strain tensor reads:
" #
← − →
− 0 αt
ε = 21 u∇0 + ∇0 u = 21 ((F − 1) + (FT − 1)) = 12 (3.84)
αt 0
vx = αY (3.85)
vy = 0
" #
1 0 α
ė = D − e · L − LT · e = 2 (3.90)
α −2α2 t
3.5 Examples 33
The deformation gradient can be decomposed via polar decomposition. The rotation tensor
and the right stretch tensor can be determined using Equations (3.56), (3.57), (3.59) and
(3.60). " #
1 2 αt
R= √ (3.91)
α2 t2 + 4 −αt 2
" #
T 1 2 αt
U=R ·F= √ (3.92)
α2 t2 + 4 αt α2 t2 + 2
E = 12 (C − 1) = 0 (3.97)
The Euler–Almansi strain tensor relates to the left Cauchy–Green stretch tensor via:
e = 12 (1 − B−1 ) = 0 (3.98)
Note that in case of a rigid rotation, the Green–Lagrange strain tensor and the Euler–Almansi
strain tensor both equal the correct value 0, where the classical strain tensor shows non-zero
components. The error made becomes larger with increasing rotation, until αt = π, half a
revolution. The velocity field for this problem is:
Ė = FT · D · F = 0 (3.104)
ė = D − e · L − LT · e = 0 (3.105)
The deformation gradient can be decomposed via polar decomposition. The rotation tensor
and the right stretch tensor can be determined using Equations (3.56), (3.57), (3.59) and
(3.60). " #
cos αt − sin αt
R= (3.106)
sin αt cos αt
" #
1 0
U= (3.107)
0 1
which is trivial for a deformation gradient that represents rotation without deformation.
Chapter 4
To define the true stress or Cauchy stress, we consider an infinitesimal surface element dS
such as given in Figure 4.1, with a unit normal vector n. An infinitesimal force dF acts on
dS:
dF = T dS = Ti ei dS (4.1)
hence, T has the dimension of a stress. The cross section areas of the surfaces perpendicular
to the xi -axes are called dSi (see Figure 4.2). It holds that:
dSi = ni dS (4.2)
35
36 Stress and Equilibrium
analogously:
dS1 = n1 dS
dS2 = n2 dS
hence:
dSi = ni dS q.e.d.
Stress components σij act on the areas dSj . The first index refers to the surface element on
which the force acts and the second index refers to the direction of the force (see Figure 4.4).
The equilibrium condition in the x1 -direction yields:
T1 dS = σ11 dS1 + σ21 dS2 + σ31 dS3 = (σ11 n1 + σ21 n2 + σ31 n3 ) dS
hence
T1 = σj1 nj
The same applies to the other directions:
T2 = σj2 nj
T3 = σj3 nj
hence:
Ti = σji nj (4.4)
4.2 Equilibrium equations 37
with S the surface of the current volume part V . Using the theorem of Gauss, the next
equation is found:
38 Stress and Equilibrium
Z
Ki = σji,j dV (4.6)
V
The force Ki must be in equilibrium with the possible volume forces (e.g. gravity force, inertia
force, induction force). The volume forces are defined as Fi per unit of volume. Now we can
write:
Z Z
σji,j dV + Fi dV = 0 (4.7)
V V
σji,j + Fi = 0 (4.8)
τ = x × T = ǫijk xj Tk (4.9)
is not equal, misses a vector
where ǫ is the permutation operator.
The total torque due to all the forces acting on the surface is represented by a surface integral
and those due to body forces are given in a volume integral. The sum of all must vanish in
equilibrium: Z Z
ǫijk xj Tk dS + ǫijk xj Fk dV = 0 (4.10)
S V
∂
Z Z Z
ǫijk xj Tk dS = ǫijk xj σmk nm dS = (ǫijk xj σmk ) dV (4.11)
∂xm
S S V
Using the chain rule to expand the differentiation and combining the two volume integrals in
one gives: Z
ǫijk δjm σmk + ǫijk xj (σmk,m + Fk ) dV = 0 (4.12)
V
The term in the parentheses vanishes when the forces in the body are in equilibrium due to
Equation (4.8). The remaining term in the integral must vanish for any arbitrary volume
which gives ǫijk σjk = 0. This results in the following three identities
proving the symmetry of the Cauchy stress tensor, i.e. σij = σji
4.3 Alternative stress representations 39
where the specific internal mechanical energy rate ẇ (per unit volume) is defined in the current
configuration as:
ẇ = σ : D (4.15)
Alternative stress measures are defined such that similar relations hold for different strain
measures. A pair of stress and strain measures for which the product results in the stress
energy rate are called work conjugate pairs.
For this reason, the Kirchhoff stress tensor τ is introduced, which is also defined on the
current configuration. This stress tensor is simply defined by
τ =Jσ (4.17)
The stress power with respect to the undeformed volume can be written as
ẇ0 = τ : D (4.18)
The Cauchy stress tensor and the Kirchhoff stress tensor are both frequently used in consti-
tutive relations.
By writing this equation in index notation, it can be proven that this can also be written as
The part between brackets is apparently work conjugated to the deformation gradient and is
called the first Piola–Kirchhoff stress P:
P = Jσ · F−T (4.21)
or
P = τ · F−T (4.22)
T
τ =P·F (4.23)
ẇ0 = P : Ḟ (4.24)
ẇ0 = P : Ḟ = (F · S) : Ḟ (4.28)
And with 2E = C − 1 = FT · F − 1:
finally, using the symmetry of the second Piola–Kirchhoff stress that easily follows from (4.25)
1
ẇ0 = S : Ė = S : Ċ (4.30)
2
This shows that the second Piola–Kirchhoff stress is work conjugated to the Green–Lagrange
strain.
Chapter 5
5.1 Introduction
Rubber and rubber-like materials remain elastic at deformations that easily exceed 100%
elongation. In Figure 5.1 e.g., the results are plotted of a tensile test on natural rubber in
which the a stretch ratio of 15 is reached. It is therefore essential that models for the me-
chanical behaviour of these elastomeric materials are valid for large deformations. The elastic
behaviour of elastomers is commonly described by so-called hyperelastic material models. In
these models a strain energy density function ψ is defined, from which the stresses can be
derived as a function of the deformation. The strain energy density represents the energy
that is stored in the material per unit undeformed volume. For elastic materials, the rate of
change of ψ is equal to the stress power ẇ0 in (4.18). The strain energy will be released if the
load is released.
16
14
12
engineering stress (Mpa)
10
0
0 2 4 6 8 10 12 14 16
engineering strain (-)
Figure 5.1: Stress–strain curve for a uniaxial tensile test on natural rubber.
41
42 Hyperelastic material models
F=R·U (5.1)
it can be concluded that ψ is only a function of the right stretch tensor U, which is symmetric.
In most models the square of U is used, referred to as the right Cauchy–Green tensor:
As shown above, the right Cauchy–Green tensor can easily be obtained without need for the
polar decomposition. Depending on the measure of deformation, one can write for the first
Piola–Kirchhoff stress P and the second Piola–Kirchhoff stress S with Equations (4.24) and
(4.30):
∂ψ
P= (5.3)
∂F
∂ψ ∂ψ
S=2 = (5.4)
∂C ∂E
The Cauchy stress can be calculated with Equation (4.27).
It should be noted that the 3 invariants are sometimes defined differently. In fact any com-
binations of the above defined invariants is by itself an invariant of C. The invariants are
strongly related to the 3 eigenvalues.
Because of the straightforward physical interpretation also the square root of I3 is often used
instead of I3 . This equals the determinant of F and represents the ratio of volumes after and
before deformation:
p V
J = I3 = det F = (5.9)
V0
We then write
ψ = ψ(I1 , I2 , J) (5.10)
∂I1 ∂(1 : C)
= =1 (5.12)
∂C ∂C
so that
∂ψ ∂ψ ∂ψ
S=2 1+2 (I1 1 − C) + JC−1 (5.16)
∂I1 ∂I2 ∂J
2 ∂ψ 2 ∂ψ ∂ψ
σ= B+ I1 B − B2 + 1 (5.17)
J ∂I1 J ∂I2 ∂J
2 ∂ψ 2 ∂ψ ∂ψ
σ= B+ I2 1 − J 2 B−1 + 1 (5.18)
J ∂I1 J ∂I2 ∂J
1
The invariants of B are equal to the invariants of C: I1 (C) = I1 (B), I2 (C) = I2 (B) and I3 (C) = I3 (B).
44 Hyperelastic material models
Compressible Neo-Hookean material model (Do we need to know this for the exam?)
G λ
(compressible neo-Hookean material model) ψ= (I1 − 3) − G ln J + (ln J)2 (5.19)
2 2
with material constants G and λ. This model only depends on the first and third invariant of
C and is known as the compressible neo-Hookean material model. In the undeformed shape
C = 1, hence I1 = 3 and J − 1 and therefore ψ = 0. The second Piola–Kirchhoff stress can
now be obtained from (5.16):
G λ
σ= (B − 1) + (ln J)1 (5.21)
J J
As expected, for an undeformed neo-Hookean material, the stress is zero.
The derivation of the material stiffness relation is out of the scope of this course, but a
quick-and-dirty evaluation at the undeformed state shows the asymptotic behavior to the
well-known small strain linear elastic stiffness. Taking the time derivative of (5.20) yields
From (3.24) and (3.39) follows that Ċ = 2Ė = 2FT · D · F. Substituting this in the previous
equation gives
The meaning of the material parameters G and λ is hereby demonstrated to be equal to the
E Eν
Lamé parameters for isotropic linear elastic material, with G = 2(1+ν) and λ = (1+ν)(1−2ν) .
For incompressible materials, the volume does not change and therefore the third invariant
of the Cauchy–Green stretch tensor I3 = J = 1. The strain energy density function is in that
case only a function of the first and second invariant:
ψ = ψ̂(I1 , I2 ) (5.26)
A problem, however, is that the stress can only be obtained up to an unknown hydrostatic
component. Or—in other words—if the deformation does not change upon a change in hy-
drostatic stress, the stress cannot be determined from the deformation only.
Although there are mathematical ways to calculate the hydrostatic stress component from
another set of equations and mathematically impose complete incompressibility, the common
solution is to allow some compressibility in the model. In fact this is physically correct.
Formally, this can be described by a general strain energy density function as in (5.10), but in
so-called nearly incompressible material models the part in the strain energy density function
that depends on the Jacobian is split-off in an additive way from the rest:
ψ = ψ̂(I1 , I2 ) + U (J) (5.27)
The second Piola–Kirchhoff stress can now be calculated from
∂ψ ∂ ψ̂ ∂U ∂J
S=2 =2 +2 volumetric strain (5.28)
∂C ∂C ∂J ∂C
A (minor) drawback of this formulation is that for a purely dilatational deformation I1 and I2
are still influenced, as can be seen in (5.6) and (5.7). Therefore, the hydrostatic stress is not
only a function of U , but also of ψ̂. It is therefore common to define a strain energy density
function that only depends on the distortional part of the right Cauchy–Green tensor:
−1/3
C̄ = I3 C = J −2/3 C (5.29)
Because an algebraic combination of invariants remains invariant, we can define distortional
invariants as
I¯1 = I1 (C̄) = J −2/3 I1 (5.30)
I¯2 = I2 (C̄) = J −4/3 I2 (5.31)
and define a—different—strain energy density function for nearly incompressible materials as
ψ = ψ̄(I¯1 , I¯2 ) + U (J) (5.32)
The model is now split in a purely distortional (isochoric) part and a purely dilatational
(volumetric) part and the second Piola–Kirchhoff stress is calculated from
∂ ψ̄ ∂U ∂J
S=2 +2 (5.33)
∂C ∂J ∂C
46 Hyperelastic material models
∂ ψ̄ ∂ I¯1 ∂ ψ̄ ∂ I¯2 ∂U ∂J
S=2 ¯ +2 ¯ +2 (5.34)
∂ I1 ∂C ∂ I2 ∂C ∂J ∂C
The derivatives with respect to C follow from:
∂ I¯1
= J −2/3 (5.37)
∂I1
∂ I¯2
= J −4/3 (5.38)
∂I2
∂ I¯1 2
= − J −5/3 I1 (5.39)
∂J 3
∂ I¯2 4
= − J −7/3 I2 (5.40)
∂J 3
and with Equations (5.12), (5.13) and (5.15) derivatives of the distortional invariants become:
∂ I¯1
1
= J −2/3 1 − I1 C−1 (5.41)
∂C 3
∂ I¯2
2
= J −4/3 I1 1 − C − I2 C−1 (5.42)
∂C 3
and
∂ ψ̄ −2/3 1 ∂ ψ̄ −4/3 2 ∂U
S=2 ¯ J 1 − I1 C −1
+2 ¯ J I1 1 − C − I2 C −1
+ JC−1 (5.43)
∂ I1 3 ∂ I2 3 ∂J
With help of the Cayley–Hamilton equation (2.84) this equation can also be written as:
∂ ψ̄ −5/3 1 ∂ ψ̄ −7/3 1 2 −1 ∂U
σ=2 ¯J B − I1 1 + 2 ¯ J I2 1 − J B + 1 (5.45)
∂ I1 3 ∂ I2 3 ∂J
5.4 Specific material models 47
∂U
= K(J − 1) (5.47)
∂J
An alternative formulation for U that was also used in (5.19) is:
1
U= K(ln J)2 (5.48)
2
For this function the derivative is:
∂U K
= ln J (5.49)
∂J J
ψ̄ = C1 (I¯1 − 3) (5.50)
The Neo-Hookean model can be obtained from statistical mechanics when vulcanized rubber
is regarded as a random network of long molecular chains, pinned at a few positions by
cross-links.
The following derivatives are required in the calculation of the stress:
∂ ψ̄
= C1 (5.51)
∂ I¯1
∂ ψ̄
=0 (5.52)
∂ I¯2
∂ ψ̄
= C1 (5.54)
∂ I¯1
∂ ψ̄
= C2 (5.55)
∂ I¯2
48 Hyperelastic material models
In order to calculate the constitutive relations the following derivatives are determined:
∂ ψ̄ 2
= C1 + 2C2 I¯1 − 3 + 3C3 I¯1 − 3
¯ (5.57)
∂ I1
∂ ψ̄
=0 (5.58)
∂ I¯2
with derivatives
n X
n
∂ ψ̄ X
¯1 − 3 i−1 I¯2 − 3 j
= = iC ij I (5.60)
∂ I¯1 i=1 j=0
n X
n
∂ ψ̄ i j−1
= jCij I¯1 − 3 I¯2 − 3
X
= (5.61)
∂ I¯2 i=0 j=1
For the determination of the material stiffness tensors, the reader is referred to e.g. [1, 3].
5.4.6 Examples
For an almost incompressible material, we can easily estimate the stress–strain response of a
uniaxial tensile test by considering the stretch ratios λ1 , λ2 and λ3 . For an incompressible
material J = √λ1 λ2 λ3 = 1 and for a uniaxial tensile test in 1-direction of an isotropic material:
λ2 = λ3 = 1/ λ1 . The deformation gradient, right and left Cauchy–Green tensor and their
inverses then become:
λ1 0 0
√
(Deformation gradient)
F= 0 1/ λ1 0 (5.62)
√
0 0 1/ λ1
and
λ21 0 0 λ−2
1 0 0
C=B=
0 1/λ1 0 C−1 = B−1 =
0 λ1 0
(5.63)
0 0 1/λ1 0 0 λ1
The value of ∂U
∂J can be derived from the boundary condition that σ2 = 0 independent of the
shape of U . We make a small error because for a particular function U , some compressibility
is allowed and λ1 λ2 λ3 is not exactly equal to one.
Substituting the obtained value for ∂U in σ1 gives:
∂J
∂ ψ̄ 2 1 ∂ ψ̄ 1
σ 1 = 2 ¯ λ1 − + 2 ¯ λ1 − 2 (5.68)
∂ I1 λ ∂ I2 λ1
The engineering stress can be derived by multiplication with the ratio of current and initial
cross section area λ2 λ3 which is here equal to 1/λ1 . Note that this is also the value of the
first Piola–Kirchhoff stress P1 .
∂ ψ̄ 1 ∂ ψ̄ 1
P1 = 2 ¯ λ1 − 2 + 2 ¯ 1 − 3 (5.69)
∂ I1 λ ∂ I2 λ1
This equation can directly be compared with the engineering stress–strain curve, derived from
experiments. The derivative with respect to λ1 is
∂ 2 ψ̄ ∂ 2 ψ̄
dP1 ∂ ψ̄ 2 2 1 ∂ ψ̄ 3 1
=2 ¯ 1+ 3 +2 λ1 − +2 ¯ +2 λ1 − 2 (5.70)
dλ1 ∂ I1 λ ∂λ1 ∂ I¯1 λ ∂ I2 λ41 ∂λ1 ∂ I¯2 λ
For the initial material stiffness (comparable to the Young’s modulus) at λ1 = 1 this yields
dP1 ∂ ψ̄ ∂ ψ̄
(λ1 = 1) = 6 + (5.71)
dλ1 ∂ I¯1 ∂ I¯2
The relation between shear modulus and Young’s modulus is 2G(1 + ν) = E, which for an
incompressible material (ν = 0.5) gives 3G = E and therefore
∂ ψ̄ ∂ ψ̄
G=2 + (5.72)
∂ I¯1 ∂ I¯2
In the generalized Rivlin model, this means that G = 2(C10 + C01 ) in order to model the
initial stiffness correct.
We can now fit the material parameters for the Neo-Hookean and Mooney–Rivlin models to
the experimental stress–strain values as presented in Figure 5.1, where we try to match the
initial shape of the curve as good as possible. The results are presented in Figure 5.2. The
parameter for the Neo-Hookean model is C1 = 0.2 and the parameters for the Mooney–Rivlin
model are C1 = 0.03 and C2 = 0.017. Only strains up to 100% are plotted. It can be
observed that both models represent the same initial stiffness and that the Mooney–Rivlin
model represents the curve well up to strains of about 50%.
50 Hyperelastic material models
0.7
experiment
Neo-Hookean
0.6 Mooney-Rivlin
engineering stress (Mpa)
0.5
0.4
0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8 1
engineering strain (-)
Figure 5.2: Stress–strain curve for a uniaxial tensile test on natural rubber.
Chapter 6
Plasticity theory
6.1 Introduction
In metals the observed inelastic deformation above a critical stress is referred to as plasticity.
Most metals have an atomic structure which consists of crystals (grains) oriented arbitrarily
(isotropic) or in a preferred way (textured) which affects the observed plastic behavior in
macroscale. Plasticity occurs by dislocation motion within this crystallographic arrangement
therefore it is expected to be isotropic or anisotropic with regard to how the grains are
oriented. In order for dislocations to move there has to be stress resolved in the crystal
structure meaning that the total strain has to be made up of an elastic part and a plastic part.
It is therefore necessary to split the total strain tensor accordingly. This split however cannot
be defined without choosing a certain direction for the strain tensors. As the dislocations
become mobilized they lead to creation of more dislocations which in turn start to hinder the
mobility of themselves. Therefore larger stresses are required as the amount of accumulated
plastic strain increases resulting in the strain hardening phenomenon.
In order to model the phenomena described very briefly above this chapter starts with a
one-dimensional formulation of the elastic-plastic split and a yield criterion. Afterwards
a general three dimensional formulation is developed by making use of Drucker’s postulate.
Commonly used yield surfaces are introduced. In order to deal with strain hardening different
formulations are given and the solution procedure is introduced.
51
52 Plasticity theory
E
σ
σy
σ
σy
−σy
σ = E εe = E (ε − εp ) (6.2)
This means that in order to find the stress, the plastic strain has to be found first.
It is clear considering the device that the stress level cannot exceed the frictional stress in
either direction. To incorporate the bidirectional response we can define a yield criterion
which describes the admissible set of stresses:
where the rate of plastic strain is defined as ε̇p = ∂εp /∂t and describes the change in the
plastic strain instead of its total value. These conditions can be described as follows: If the
absolute value of the stress is less than the yield stress there is no change in the plastic strain;
If the stress is equal to the yield stress on the positive side, there will be a positive change in
the plastic strain; If the stress is equal to the yield stress on the negative side there will be a
negative change in the plastic strain.
Note that conditions 2 and 3 can actually be represented by a single equation:
λ̇ ≥ 0
φ(σ) ≤ 0 (Kuhn-Tucker conditions)
λ̇ φ(σ) = 0 (6.5)
Equation (6.7) reveals that once the yield stress is reached all strain increment becomes plastic
which is the solution that we are after.
must be defined to account for arbitrary stress states and the direction of the plastic strain rate
must be postulated in order to find a solution. Other considerations such as the Kuhn-Tucker
and the consistency conditions can be generalized to multi-axial form in a straightforward
manner by replacing the scalar components with tensorial counterparts.
s a s b
e e
Figure 6.4: Two instable stress-strain curves
I
σ : dε ≥ 0 (6.8)
dσ : dε ≥ 0 (6.9)
dσ : dεp ≥ 0 (6.10)
In words what (6.10) says is that the plastic strain increment cannot oppose the stress incre-
ment. It should be noted that this postulate applies generally for metals but other materials
for instance soil do not have to possess this property.
The consequences of Drucker’s postulate are very significant in plasticity theory. Consider
for instance the arbitrary yield surface once again with a closed stress path starting from σ a
56 Plasticity theory
onto the yield surface σ y to an increment of ∆σ on the yield surface (consider hardening)
and then back to σ a as shown in figure 6.6 which would result in a one-dimensional results
as shown in figure 6.5.
The mechanical work done by an external agency in this cycle according to Figure 6.6 is
calculated as:
I I
(σ − σ a ) : dε = (σ − σ a ) : d(εe + εp ) ≥ 0 (6.11)
The elastic part of the integral can be shown to be zero at a closed stress cycle. Since σ : εe
gives the elastic strain energy it vanishes in a cycle. Furthermore σ a is constant therefore
what remains in the integral is only dεe which also vanishes in a closed stress cycle. WHat
remains is the plastic part which only exists in the path traveling with the yield surface by
∆σ. This part can be linearized for small increments and using trapezoidal rule the integral
can be written as:
1
2 [(σ y − σ a ) + (σ y + ∆σ − σ a )] : dεp = (σ y − σ a ) : dεp + ∆σ : dεp (6.13)
Note that for a hardening material the last term is always positive (equation (6.10)). The
remaining term is the important result that we are after:
(σ y − σ a ) : dεp ≥ 0 (6.14)
The consequences of (6.14) are twofold. First, let’s consider the direct result the double
contraction operation. The difference of the stress at the yield surface and an arbitrary stress
inside is projected on the plastic strain increment resulting in a positive value. This can only
happen if the projection is done with a sharp angle (−π/2 ≤ α ≤ π/2). This is explained
further with the help of Figure 6.7, in which the vectors σ a and σ ∗a are arbitrary vectors.
On the figure the dashed areas are the admissible directions for the plastic strain increment
due to Drucker’s postulate given in equation (6.14). But at the current state of deformation
which happens on the yield surface the previous stress (σa ) is irrelevant, the postulate should
6.3 Plastic deformation at multi-axial stress states 57
hold for any admissible stress. The only direction of plastic strain that allows this is the
normal to the yield surface as shown in figure 6.8. This result is known as the principle of
normality.
In case dεp is not perpendicular to the yield surface, an arbitrary σa would exist for which
the postulate is violated:
∂φ
It can be shown that ∇φ = ∂σ with two examples. First, assume a circular function
∂φ
2 2 2
φ = x + y − R . The gradient of this function is ∂x = ( ∂φ ∂φ
∂x , ∂y ) = (2x, 2y). The latter
vector is perpendicular to the circular function φ.
Second, assume a vertical line at a distance c from the origin φ = x − c = 0. The gradient of
∂φ
this function is ∂x = ( ∂φ ∂φ
∂x , ∂y ) = (1, 0). The latter vector is perpendicular to the vertical line.
The principle of normality can be used to derive the demand for convexity of the yield surface.
For a visual explanation refer to figure 6.10. Equation (6.14) must be valid for any arbitrary
stress in the elastic domain. If the yield surface is non-convex though this might conflict
with the normality rule where the resulting stress increment would be instable for a plastic
deformation direction that is normal to the yield surface. As a result it is required that the
yield surface is a convex function.
Note that this only applies to regular points. For singular points, in case of an intersection of
two yield surfaces, see Figure 6.11, the convexity of the yield surface is guaranteed if φ1 = 0
6.4 Elastic - ideal plastic constitutive equations 59
∂φ1 ∂φ2
dεp = dλ1 + dλ2 (6.17)
∂σ ∂σ
∂φ
ε̇p = λ̇ (6.20)
∂σ
Substitution of equation (6.20) in (6.19) gives:
∂φ
σ̇ = E : ε̇ − λ̇ (6.21)
∂σ
Kuhn-Tucker conditions can be written in tensorial form as follows:
λ̇ ≥ 0
φ(σ) ≤ 0
λ̇ φ(σ) = 0 (6.22)
60 Plasticity theory
∂φ ∂φ
E: :E
σ̇ = E −
∂σ ∂σ
: ε̇ = (E − Y) : ε̇ (6.26)
∂φ ∂φ
:E:
∂σ ∂σ
• the yield surface reduces to the yield stress in case of a uniaxial stress state,
For metals the following are also observed in experiments: hydrostatic pressure does not cause
plastic deformation and the plastic deformation is incompressible. The first condition implies
that the yield surface does not depend on pressure. The second condition can be expressed
as:
εpii = 0 =⇒ ε̇pii = 0
With (6.16) follows:
∂φ
ε̇p = λ̇ (6.27)
∂σ
In general λ̇ 6= 0 , which gives:
trace
∂φ
=0 (6.28)
∂σii
6.5 Multi-axial yield functions 61
multi-axial stress state means that the diameter of the largest circle of Mohr equals the yield
stress. The diameters of both other circles have to be smaller or equal to the yield stress. In
other words, the Tresca yield criterion is nothing more than:
|τmax | = 12 σy (6.29)
In three dimensions, the Tresca yield criterion has the following form:
φ = (σ1 − σ2 )2 − σy2 (σ2 − σ3 )2 − σy2 (σ3 − σ1 )2 − σy2 = 0
(6.30)
With this equation it is easy to prove that the Tresca yield criterion does not depend on the
pressure. Since the pressure does not influence the yield criterion, the yield criterion is an
62 Plasticity theory
The cross section of the plane perpendicular to the {1,1,1,}-direction is always identical. The
plane σ1 + σ2 + σ3 = 0 is also known as the π-plane. The cross section of the π-plane with
the Tresca yield surface is represented in figure 6.13.
The relation between deviatoric stresses and the total stresses with respect to an arbitrary
reference frame, is given by:
in which p = 13 σkk = 13 (σ11 + σ22 + σ33 ). The elastic boundary can be represented by:
The term sij sij is invariant, it does not change if the tensor components are transformed to
another reference frame (because it is a scalar quantity). Therefore, Equation (6.35) holds for
every arbitrary Cartesian coordinate system. The equivalent stress according to Von Mises is
defined by:
q
σ = 32 sij sij (6.36)
The factor 32 is included in order to achieve that in a uni-axial tensile test the uni-axial stress is
equal to the equivalent stress (see section 6.5.6). This section also shows the relation between
k and σy . Hence, the Von Mises yield surface in terms of deviatoric stresses is defined as:
φ = 23 s : s − σy2 = 0 (6.37)
In case of using stresses instead of deviatoric stresses, the Von Mises yield surface reads,
according Huber and Hencky:
The Von Mises yield surface is also pressure independent, and as a result, will be a prismatic
in the three dimensional stress space. Writing the Huber and Hencky equation in principal
stresses, the equation describes a cylinder with its axis in the {1,1,1}-direction. The cross
section of the yield surface with the σ1 − σ2 - plane is elliptical, the cross section with the
π-plane is a circle, see Figure 6.14. In three dimensions, the Von Mises yield surface is
represented in Figure 6.15. Besides, it is simple to prove that the Von Mises yield surface
fulfills all demands, as described in the introduction.
s1
von Mises
p-plane
s2 s3
Tresca
s2 s3
Figure 6.16: The Von Mises criterion, bounded by the lower bound and the upper bound
disadvantage of the Tresca yield criterion compared to the Von Mises yield criterion, is that
it contains singular points.
principal stress space, the Hill yield criterion is then represented as follows:
σ2
σ1
R=1/2
R=1
R=2
R=3
ε3 is the strain in thickness direction. The Hill yield criterion degenerates to the Von Mises
yield criterion if R equals 1 (isotropic material behavior). Besides, it is simple to prove that
the Hill yield surface fulfills all demands, as described in the introduction. Examples of the
Hill yield surface for various R-values are given in Figure 6.17.
Ideally, the material is assumed incompressible. In two dimensions, the material behavior can
be characterized by a material which shows Coulomb friction:
The stress-strain relation is:
plastic deformation rate has a component in positive direction, which also means that the
plastic deformation rate will be positive and hence, that the volume will increase. In other
words, applying a pressure should lead to an increase in volume, which is strongly arguable.
Therefore plastic deformation is only allowed to occur in τ -direction, see Figure 6.19.
In general, for this kind of material behavior, it is assumed that the plastic deformation rate
is determined by a potential which not equals the yield surface (this in contrast with the
former material behavior where both coincide). This is called non-associated flow.
sij sij − 2k 2 ≤ 0
sij = 0 if i 6= j
Furthermore:
sij sij = s211 + (− 12 )2 s211 + (− 12 )2 s211 = 32 ( 23 σ11 )2
in other words:
sij sij = 2k2 = 23 σy2 (6.44)
hence:
σy
k=√ (6.45)
3
The yield surface is obtained as in equation (6.20) using the above:
φ = 32 s : s − σy2 ≤ 0 (6.46)
Often an alternative version of the yield surface is used based on the equivalent stress concept:
VM
φ∗ = σeq − σy ≤ 0 (6.47)
q
with VM
σeq = 32 s : s.
With this result at hand now the plastic strain can be related using normality to the yield
surface using the previously obtained result:
∂φ
ε̇p = λ̇
∂σ
and consequently using (6.46):
ε̇p = 3 λ̇ s (6.48)
Using the consistency condition λ̇ can be determined as:
1 s : E : ε̇
λ̇ = (6.49)
3s:E:s
Based on the isotropic elasticity assumption this result can be further reduced to:
1 n
λ̇ = √ : ε̇ (6.50)
3 s:s
√
where n is the direction of the deviatoric stress: n = s/ s : s.
68 Plasticity theory
E : ss : E
σ̇ = E− : ε̇ = (E − 2G nn) : ε̇ (6.51)
s:E:s
In this case, the deformation history is given by the logarithm of the strain ε. By introducing
a scalar deformation value, the equivalent plastic strain, this equation can be generalized to
an arbitrary deformation state (like the equivalent stress)
Zt q
2 p p
ε= 3 ε̇ij ε̇ij dt (6.52)
t0
Note that the equivalent plastic strain cannot decrease. The factor 23 is introduced to make
the equivalent plastic strain equal to the logarithmic strain in case of a uni-axial tensile test
(ε̇11 = ε̇11 , ε̇22 = ε̇33 = − 21 ε̇11 ) . Proof:
Zt q Zt q Zt q Zt
2 2 2 2 1
ε= 3 ε̇ij ε̇ij dt = 3 (ε̇11 + ε̇222 + ε̇233 ) dt = 3 (1 + 4 + 14 )ε̇211 dt = ε̇11 dt = ε
t0 t0 t0 t0
(6.53)
∂φ ∂φ
φ̇ = σ̇ij + β̇ij = 0 (6.55)
∂σij ∂βij
λ̇ = 0
as well as:
β̇ij = 0
It is assumed that the relation between β̇ and ε̇p can be written as:
∂φ
β̇ij = f (σij , βij , T, . . .)λ̇ (6.59)
∂βij
Now equation 6.55 can be written as:
∂φ
φ̇ = σ̇ij + f λ̇ = 0 (6.60)
∂σij
In the subsequent subsections the function f will be derived for both isotropic and kinematic
hardening. Substitution of (6.21) in (6.60) yields the next expression:
∂φ ∂φ
φ̇ = Eijkl ε̇kl − λ̇ + f λ̇ = 0 (6.61)
∂σij ∂σkl
which gives the following expression for λ̇:
∂φ
Eijkl ε̇kl
∂σij
λ̇ = (6.62)
∂φ ∂φ
Epqrs −f
∂σpq ∂σrs
The general constitutive equations for elastic non-ideal plastic material behavior can be de-
rived by combining equations (6.62) and (6.21).
∂φ ∂φ
Eijmn Etukl
∂σmn ∂σtu
σ̇ij = E
ijkl − ε̇kl (6.63)
∂φ ∂φ
Epqrs −f
∂σpq ∂σrs
Or in tensor notation:
∂φ ∂φ
E:
:E
σ̇ = E −
∂σ ∂σ : ε̇
(6.64)
∂φ ∂φ
:E: −f
∂σ ∂σ
A brief version of the expression is:
σ̇ = (E − Y ∗ ) : ε̇ = (E − (1 − h)Y) : ε̇ (6.65)
with:
∂φ ∂φ
E: :E f
Y= ∂σ ∂σ and h =
∂φ ∂φ ∂φ ∂φ
:E: f− :E:
∂σ ∂σ ∂σ ∂σ
with h a scalar which depends on the hardening. The yield behavior is expressed by the
tensor Y in (6.65).
70 Plasticity theory
6.7 Hardening
6.7.1 Introduction
Previously a constitutive equation was derived for elastic ideal plastic material behavior.
Since hardening does not occur, the yield surface will not change. In practice, most materials
show hardening. As a result the yield surface will change in size or position in the stress
space, see Figure 6.20. Two extreme types of hardening can be distinguished:
1. Isotropic hardening The shape and position of the yield surface remain the same, the
size of the yield surface uniformly increases, see Figure 6.21.
2. Kinematic hardening The shape and the size of the yield surface remain the same, the
position of the yield surface shifts in the stress space, see Figure 6.22.
The difference between both types of hardening can also be clearly spotted when looking at
the stress-strain curve, see Figure 6.23. In case of kinematic hardening the elastic part during
unloading equals the initial elastic part. In case of isotropic hardening, the elastic part has
increased during unloading. The hardening of a material is not purely kinematic or isotropic.
Some parts of the material behavior are better described with the isotropic hardening and
6.7 Hardening 71
vice versa, see last subsection on hardening. First, a constitutive equation will be derived to
take into account hardening. Then the general form of this equation will be applied to both
types of hardening.
It is assumed that the yield stress is not a constant, but depends on the equivalent plastic
strain (which is a measure for the plastic strain). Note: in the general form (6.54), β is
defined as the hardening tensor. Equation (6.66) is a special case of (6.54) when it is assumed
that βij = εp δij . The relation between the yield stress and the equivalent plastic strain is
p
determined by a tensile test. During this tensile test, the stress σ11 and the strain ε11 (so also
p
the plastic strain ε11 ) in the tensile direction are measured. It is defined that the equivalent
plastic strain equals the plastic strain in the tensile direction εp = εp11 . To determine the
unknown function f in equation (6.60), the derivative of the yield surface is taken (6.66):
dσy p
φ̇ = 3sij ṡij − 2σy ε̇ = 0 (6.67)
dεp
The first term in (6.60) equals the first term in (6.67) since in case of Von Mises yield behavior:
∂φ
σ̇ij = 3sij σ̇ij = 3sij ṡij (6.68)
∂σij
dσy p
f λ̇ = −2σy ε̇ (6.70)
dεp
This expression can be elaborated further, using the relation between ε̇p and λ̇, equation
(6.20):
ε̇pij = 3λ̇sij = µ̇sij (6.71)
First a relation between the equivalent plastic strain rate ε̇p and the plastic strain rate tensor
ε̇pij has to be determined. This relation is defined by the tensile test. The plastic volume
changes are zero, see subsection 6.5. So, in case of a tensile test, the following relations can
be determined:
ε̇p22 = ε̇p33 = − 12 ε̇p11 (6.72)
The following norm is used to define the equivalent plastic strain rate:
q
ε̇p = const ε̇pij = const ε̇pij ε̇pij
6.7 Hardening 73
Filling in the Prandtl-Reusz equation (6.71) gives a relation between ε̇p and λ̇.
q
ε̇p = 3λ̇ 23 sij sij (6.74)
Again, the expression under the square root is found, using the tensile test. The deviatoric
stress components are written as:
s11 = σ11 − 31 σ11 = 23 σ11 = 32 σy (6.75)
s22 = s33 = − 13 σ11 = − 31 σy (6.76)
Filling in the above stress components in (6.74) gives:
ε̇p = 2λ̇σy (6.77)
The expression for f in case of isotropic hardening, using the Von Mises yield criterion, is
found by substituting equation (6.77) in (6.70):
dσy
f = −4σy2 (6.78)
dεp
Finally, note that the equivalent plastic strain is determined as follows (and not calculated
directly from the plastic strains):
Zt1 Zt1 q
p p 2 p p
ε = ε̇ dt = 3 ε̇ij ε̇ij dt (6.79)
t0 t0
This strategy is followed since the total strain can be zero during cyclic loading, while the
equivalent plastic strain can not decrease by definition.
Nadai model
For many steels, a good approximation of a large part of the true stress–strain curve can be
obtained with a power law, e.g. the Ludwik–Nadai or Nadai relation1
σy = Cεn (6.80)
C is the coefficient of hardening and n is the exponent of hardening.ε should formally be the
plastic strain in the tensile direction, but for large deformations the elastic part is usually
neglected. To determine the yield stress according to the Nadai hardening law, the following
equation is used:2
σy = σ0 + C (εeq + ε0 )n (6.81)
with σ0 the offset yield stress, ε0 the offset plastic strain, and C and n are power law variables.
1
Also called the ‘Hollomon’ relation.
2
Also called the ‘Swift’ relation.
74 Plasticity theory
Voce model
If the hardening rate is plotted as a function of the stress, often a linearly decreasing relation
can be observed. This part of the hardening curve is often denoted as Stage III hardening.
A relation that represents this behaviour is as follows:
dσf σ
θ= = θ0 1 − sat (6.82)
dε σ
where σ sat is the saturation stress. The resulting flow stress as a function of the strain is
written as
ε
σy = σy0 + ∆σ 1 − exp − (6.83)
ε0
where ∆σ = σ sat − σy0 and ε0 = σ sat /θ0 . This model is known as the Voce model.
The Voce hardening law is evaluated as:
(
σy = σy0 if εeq ≤ ε1
εeq −ε1 (6.84)
σy0 + ∆σ 1 − exp − ε0 if εeq > ε1
with σy0 the initial yield stress and ∆σ the increase of the yield stress until saturation is
reached.ε0 is the plastic strain after the yield point plateau was reached; the plastic strain
parameter ε1 indicates the length of the yield plateau.
Some metals with small grains, at temperatures above 0.5 Tm and at specific strain rates
exhibit so-called super-plastic behavior. Under these conditions they can reach engineering
strains up to 1000 %. E.g. an aluminum alloy 2004, when deformed at 430 ◦ C and a strain
rate of 2 × 10−3 s−1 , gives et = 800 % at a flow stress of 10 MPa. This same alloy when heat
treated to the T6 condition has a tensile strength of 450 MPa and is used for the manufacture
of a range of aircraft components [8, 127].
The superplastic deformation behavior is often described by the stress–strain-rate relation
σ = kε̇m (6.85)
and, in the superplastic range, 0.5 < m < 1. The parameter m is itself a function of the
strain rate (among others) and it is the task of the manufacturer of products to control the
process as to assure that the value of m is near a maximum. For high values of m localization
is postponed and high strains can be reached.
Combining, for lower temperatures, the strain hardening and strain-rate hardening equations,
we get
σ = Kεn ε̇m (6.86)
In physically based hardening models, the evolution of the flow stress is predicted by consider-
ing the physical mechanisms of plastic deformation. Ideally, the influence of work hardening,
strain rate and temperature, and their interactions, are included.
Plastic deformation in f.c.c. and b.c.c. metals is mainly the result of dislocation movement on
the active slip systems. Hence, the resistance to dislocation movement determines the yield
stress. In a physically based model the yield stress is often decomposed into a strain and
strain rate independent stress σ0 , a dynamic stress σ ∗ that depends on the strain rate and
temperature and a term σw that incorporates the work hardening:
The dynamic stress σ ∗ is the driving force that is required to move dislocations through the
atomic lattice at a certain velocity. The velocity of a dislocation depends on the rate at
which atoms change position in the atomic lattice. This rate is described with a probabilistic
approach in statistical thermodynamics. An approximation of the resulting implicit relation
76 Plasticity theory
for σ ∗ for relatively low temperatures yields the description of Krabiell and Dahl [6] that is
often used in simulation programs:
0 if ε̇ < ε̇0 exp(−∆G0 /kT )
p
∗ ∗ kT ε̇
σ (ε̇, T ) = σ0 1 + ∆G0 ln ε̇0 if ε̇0 exp(−∆G0 /kT ) < ε̇ < ε̇0 (6.89)
σ ∗ if ε̇ > ε̇
0 0
The work hardening part of the model, σw , takes the evolution of the micro-structure into
account. A relatively simple one-parameter model only considers the dislocation density ρ
to be responsible for the hardening. The relation between the dislocation density and σw is
given by the Taylor equation:
√
σw = αGb ρ (6.90)
where α is a scaling parameter of order 1 and G is the shear modulus. This equation is sup-
ported by a dimensional analysis and numerous experiments and is used in a large number of
physically based models. The essential part in these models is the evolution of the dislocation
density ρ. Several models are available, and the reader is referred to literature.
σy dκ = σ · dε (6.91)
For a uniaxial loading process κ is taken equivalent to the uniaxial plastic strain and con-
sequently, the uniaxial stress σun is equal to the flow stress σf . The hardening found in a
uniaxial test is presented as the uniaxial stress σun as function of the uniaxial strain εun .
The equivalent stress σy in the numerical description of the yield functions is now considered
to vary in the same way depending on κ. For proportional loading paths, the ratio between
components in the stress tensor remains constant (α = σ2 /σ1 ), as well as the ratio between
components in the strain tensor (β = ε2 /ε1 ). From this consideration, relations can be de-
rived for the expression of any component of the stress tensor as function of any component
of the strain tensor. For the Nadai relation this yields simply:
where
C ′ = Cf (β, n) (6.93)
6.7 Hardening 77
moved center):
2
φ = 32 (sij − αij )(sij − αij ) − σy,0 =0 (6.94)
Now the yield stress is constant since the size of the yield surface does not change. To
determine the unknown in equation (6.60), equation (6.94) is differentiated first.
The displacement of the yield surface in the stress state has the same direction as the plastic
deformation. The first term in (6.60) equals the first term in (6.95), using the Von Mises
yield surface and the assumption of Prager (6.96):
∂φ
σ̇ij = 3(sij − αij )σ̇ij = 3(sij − αij )ṡij (6.97)
∂σij
A comparison of the second term in (6.60) and the second term in (6.95) gives an expression
for f :
f λ̇ = −3(sij − αij )α̇ij (6.98)
Using the assumption of Prager (6.96) and the principle of normality gives:
∂φ
f λ̇ = −3(sij − αij )h∗ λ̇
∂σij
78 Plasticity theory
The derivative of the yield surface to the stress is given in (6.97). This gives:
2 dσy
h∗ = 3 (6.102)
dεp
2 dσy
f = −4σy,0 (6.103)
dεp
In case of a bilinear approximation of the stress-strain curve, as given in Figure 6.26, the
factor h∗ is written as:
h∗ = 23 Et (6.104)
The model described above is a linear hardening model. The evolution of the backstress,
Equation (6.96), can be written in incremental form as:
∂φ
∆α = ∆λh∗ (6.105)
∂σ
Another well known kinematic hardening model is the Armstrong–Frederick hardening
model. The evolution of the backstress is described by:
∗ ∂φ
∆α = ∆λ h − hl α (6.106)
∂σ
The second term in the brackets gives a smooth transition from the elastic to the plastic
region.
6.8 Elastic-viscoplastic material 79
6.8 Elastic-viscoplastic material (Do we have to know this for the exam?)
6.8.1 Introduction
Creep is the occurrence of plastic deformation during prolonged constant loading. On the
other hand, relaxation means that at a constant deformation, the stress decreases with in-
creasing time. Shape recovery after load removal is related to both phenomena. Finally, there
is a relation with strain-rate dependent material behavior. The material behavior of metals
is almost independent of the strain-rate, when this is not too large. Therefore metals do have
a so called static stress-strain curve. This behavior can be described with a Bingham model
with a spring in series. However, plastic strain occurs during elongated loading. This behavior
can be modeled with a Maxwell model (area of secondary creep); this model does not describe
the plastic behavior for stresses higher than the yield stress. In case of large strain rates, the
influence of the strain rate on the material behavior of metals can not be neglected, like in the
rolling process. An increase of the strain rate yields an increase of the stress. For each fixed
strain rate, the stress-strain curve is higher than the static stress-strain curve, but with the
same shape. Polymers are sensitive for creep and are strain rate dependent for all strain rates.
Polymers can be described with the same model in the both the creep area as in the plastic
deformation area. The model looks like the Maxwell model, but has a non-linear damper. In
general, the logarithm of the strain is proportional with the stress. Strain rate dependency
in metals and polymers and creep in polymers is called viscoplastic material behavior and is
treated in subsection 6.8.3. Subsection 6.8.2 treats creep in metals.
be as constant as possible since a small change in temperature can have a large effect on the
results. The next areas can be distinguished in Figure 6.27:
The creep curves depend on the temperature and the stress. Creep is a time dependent
phenomenon; plastic deformation occurs while the yield stress is not exceeded. This is called
viscoelastic material behavior. When the yield stress is exceeded, the material behavior can
become both time and history dependent: plastic deformation is caused by plasticity and
creep. This phenomenon is called viscoplastic material behavior. We restrict ourselves to
the area with the constant strain rate, secondary creep. Starting point will be the Maxwell
model, see Figure 6.28.
Besides, isotropic material behavior is assumed and it is assumed that the relation is also
valid for other stress states. The effective stress becomes then:
q 1
σeq = 32 (s : s) 2 (6.109)
ε̇vp = λs (6.111)
To obtain a stress strain relation, Hooke’s law is used in combination with equation 6.107:
σ = E : εe = E : (ε − εvp ) (6.113)
Since we only have an expression for the viscoplastic strain rate, the derivative of the above
equation is taken:
σ̇ = E : (ε̇ − ε̇vp ) (6.114)
substitution of equations 6.111 and 6.112 in 6.114 yields and expression for the viscoplastic
stress state:
ε̇vp
σ̇ = E : (ε̇tot − 2 s) (6.115)
3 σeq
Note that, contrary to the elastic-plastic model, the equivalent viscoplastic strain rate does
not have to be eliminated since the value is known through equation 6.110
Chapter 7
7.1 Introduction
Currently the most common numerical method that is used for solving mechanics problems
involving nonlinearity in material behavior is the Finite Element Method. In previous courses
this method was introduced (Introduction to the Finite Element Method) and further elabo-
rated (Numerical Methods in Mechanical Engineering). In the following, first, a brief summary
of the method is supplied with a focus on material nonlinearities. Afterwards the necessary
integration algorithms for the constitutive models that were introduced in Chapter 6 will be
given.
83
84 Finite Element Method
The first integral is nothing else than the contribution of surface forces Ti . In the second
integral, only the symmetric part (δDij ) of δvi,j contributes to work due to the required
symmetry of Cauchy stress:
δDij = 12 (δvi,j + δvj,i ) (7.5)
Equation (7.6) is usually referred to as the weak form of the equilibrium equation. However
there are alternative ways it can be interpreted. If the chosen arbitrary weight functions,
δvi , are considered to be virtual, kinematically admissible infinitesimal displacements then
the resulting equation is an expression of the virtual work theorem. In this case the first
term represents the internal virtual work and the last two represent the virtual work done
by external forces. And since the volume change would also be infinitesimal in this case this
equation can be solved using the linear finite element method.
In case of large deformations it is valuable to consider the weight functions, δvi , as kinemati-
cally admissible virtual velocities that are not necessarily infinitesimal. In this case equation
(7.6) represents the virtual power. It also implies that the derivatives are with respect to
current coordinates and that the volume is not constant. Accordingly, the tensor δDij can be
interpreted as the virtual rate of deformation.
In tensor notation equation (7.6) is given as:
Z Z Z
δD : σ dV = δv · F dV + δv · T dS ∀δv (7.7)
V V S
n
X
ξ= N α ξα (7.9)
α=1
7.4 Iterative Newton–Raphson procedure 85
where N α is the time independent interpolation function for node α of the element with n
the number of nodes per element and ξ represents any independent field to be interpolated in
space. In the following the summation operator will be omitted and the summation convention
will be assumed.
In order to discretize the weak mechanical equilibrium, all continuous field variables and their
gradients are determined by the interpolation functions and the nodal values, leading to:
or:
v v 1 v 2 α
1 1 1 v1
v2 = N1 v21 + N2 v22 + . . . + Nα v2α (7.11)
v v1
v2
vα
3 3 3 3
δv = N α δvα (7.12)
δDij = 12 (δvi,j + δvj,i ) = 12 N,jα δviα + N,iα δvjα = δvkα 21 δki N,jα + δkj N,iα = δvkα Bkij
α
(7.13)
δD = δvα · Bα (7.14)
where Bα is a third order tensor that relates the rate of deformation tensor to the nodal
velocity vector. Using Equation (7.12) and (7.14) in Equation (7.7) gives the discretized form
of the virtual work (power):
Z Z Z
δvα · Bα : σ dV = δvα · N α F dV + N α T dS
V V S
α
⇒ δv · Fαint = δv α α
· Fext (7.15)
where Fαint is the internal force vector and Fαext the external force vector. Since Equation
(7.15) must hold for all kinematically admissible virtual velocity fields, the following equality
is obtained for the internal and external force vector:
Repeatedly solving the linearized set of equations can lead to an approximation of the desired
solution, provided a converging process.
The starting point of the Newton–Raphson procedure is a displacement vector ∆u(0) . Next,
define:
∆u = ∆u(0) + ∆∆u (7.17)
Consequently, Equation (7.16) can be written as:
Note that the superscript α, indicating the nodes, is omitted in this section to enhance
readability. Applying a Taylor series expansion to this equation yields:
∂Fint
Fint (∆u) = Fint (∆u(0) ) + | (0) · ∆∆u + O(∆∆u2 ) = Fext (7.19)
∂u ∆u
Neglecting the higher order terms O(∆∆u2 ) and rearranging gives:
∂Fint
K(0) · ∆∆u(1) = Fext − Fint (∆u(0) ) = R(0) with K(0) = | (0) (7.20)
∂u ∆u
where K(0) is the tangent stiffness matrix and R(0) is the residual force vector which van-
ishes when the exact solution is found. The unknown displacement increment ∆∆u(1) can be
solved using Equation (7.20). However, ∆u(0) + ∆∆u(1) is in general not equal to the exact
solution, since the higher order terms are omitted. It is likely that the new approximation
of the displacement vector, ∆u(1) = ∆u(0) + ∆∆u(1) , is a better approximation of the exact
solution than the old approximation ∆u(0) . The procedure is repeated again using ∆u(1) as
a new initial approximation. With the new approximation ∆u(1) , a new stiffness matrix K(1)
and internal force vector Fint (∆u(1) ) are determined. Subsequently, ∆∆u(2) is calculated
using Equation (7.20) (with K(1) and Fint (∆u(1) )) and the new displacement vector is set to
∆u(2) = ∆u(1) + ∆∆u(2) . This procedure is repeated until a user-defined accuracy is reached,
see algorithm below:
initialize
∆u(0) = 0
R(0) = Fext − Fint (∆u(0) )
k=0
WHILE (kR(k) k > tolerance) DO
determine K(k)
K(k) · ∆∆u(k+1) = R(k)
∆u(k+1) = ∆u(k) + ∆∆u(k+1)
determine Fint (∆u(k+1) )
R(k+1) = Fext − Fint (∆u(k+1) )
k := k + 1
END
In the above algorithm, it is assumed that Fext does not depend on ∆∆u. However, if Fext
7.4 Iterative Newton–Raphson procedure 87
depends on ∆∆u (e.g. in pressure loads), this can easily be incorporated in the algorithm.
In one dimension, the iterative Newton–Raphson process can be graphically represented as
shown in Figure 7.1. The exact solution is represented by û in this figure.
To avoid the creation of a new system of equations, the modified Newton–Raphson method
can be applied. In this method, the stiffness matrix in iterate k equals the stiffness matrix in
the first iteration (K(k) = K(0) ). While the convergence of this method will be slower, time
is saved by not computing the tangent stiffness matrices within each iteration, which in the
end can lead to a more economical method.
In general, the Newton–Raphson method converges, provided that the initial guess is close
to the exact solution. Divergence can occur for a too distant initial guess, see Figure 7.2.
If convergence cannot be obtained for a certain external force, this force can be subdivided
in several smaller loadsteps, and subsequently solving the system of equations for each of
the load substeps. The internal force vector Fint (∆u(k) ) and the stiffness matrix K(k) must
To determine the internal force vector, according Equation (7.15), the stress at the end of
the time increment must be calculated. This is done by a stress update algorithm, which is
explained in Section 7.6. The derivation of the stiffness matrix is given in Section 7.5.
Upon deformation, a start geometry (at time = t(n) and a final geometry (at time = t(n+1) ) can
be distinguished. Hence, the integral to compute the internal force vector can be evaluated
on the start geometry or on the final geometry. The geometrically linear equations are based
on the starting geometry whereas the geometrically nonlinear equations are based on the
final geometry. Since by definition equilibrium has to be fulfilled on the final geometry, the
geometrically nonlinear case represents a more realistic solution.
On the other hand for some applications it can be the case that nonlinear regime in the
material is reached before geometrical nonlinearities become significant, especially when an
incremental scheme is used with small time steps where rotations are also negligible. In the
following sections both cases are treated.
The derivative of the internal force vector reads (see, Equation (7.15)):
Z
Ḟαint = Bα0 : σ̇ dV0 (7.23)
V0
where the subscript 0 refers to the start geometry. According to Equations (6.26) or (6.65)
the constitutive equation can be written as:
σ̇ = Ly : ε̇ (7.24)
7.5 Determination of the stiffness matrix 89
where J is the Jacobian (determinant of the deformation gradient F). The time derivative of
the internal force vector becomes:
→ α
− →
−
Z Z
α −T −T
Ḟint = 1∇0 N : σ̇ · F J dV0 + 1∇0 N α : σ · Ḟ J dV0 (7.29)
V0 V0
where the time derivative of J is neglected since the volume changes are only elastic and
very small. Again without proof, it is stated that when the time derivative of the density is
incorporated in the constitutive equations, the time derivative of J cancels out with the extra
term concerning the density.
−T
An expression must be found for the time derivative of Ḟ :
0 = 1̇ = (F · F−1 )˙ = Ḟ · F−1 + F · Ḟ−1 ⇔
−1 −1 −1 −1 −1
Ḟ = −F · Ḟ · F = −F ·L·F·F = −F−1 · L ⇒
−T T −T
Ḟ = −L · F (7.30)
90 Finite Element Method
The transformation of the time derivative of the internal force vector from the initial to the
current configuration is obtained through Equation (7.28):
→
− α →
−
Z Z
α T
(current configuration) Ḟint = 1∇N : σ̇ dV − 1∇N α : σ · L dV (7.32)
V V
Next, the derivative of the Cauchy stress must be defined. In the geometrically nonlinear case
we have to account for large changes in geometry. First, a situation is considered in which
the strain remains elastic but the displacements and rotations are large. In that case Hooke’s
law still applies:
σ = E : εe (7.33)
Taking the material derivative gives:
σ̇ = E : ε̇e (7.34)
However, from chapter 3, it is known that the classical strain tensor is sensitive to rotations;
the strain tensor does not remain 0 upon rotation. In chapter 3 it is shown that the Green
Lagrange strain tensor is insensitive for rotations. However, if ε̇ is replaced by Ėgl in the
above equation, then the equation is not valid anymore in case of rotations. This can be
shown with the next example.
Assume a bar placed along the x-direction and load it with a small tensile stress σ 0 = σx0 .
Then the bar is rotated until it is aligned with the y-axis, without applying any further
deformation, i.e. Ėgl = 0. However, the stress components do change since these are related
to the x- and y-direction, see Figure 7.3. This means that, if we want to replace the strain
rate tensor with the Green Lagrange strain tensor, we have to define a stress measure which
is also independent on rotation. The Cauchy stress can be expressed in components referring
either to the global base vectors ei or local, co-rotating, base vectors gi :
with τ = ei τij ej . The stress tensor τ is invariant under superimposed rigid rotation. Hence,
the derivative of Hooke’s law in invariant measures reads:
τ̇ = E : Ėgl (7.37)
with E the elasticity tensor and Ėgl the derivative of the Green Lagrange strain, as defined
by Equation (3.39):
T
Ėgl = F · D · F (7.38)
or, using polar decomposition and assuming small elastic deformations (U = I):
T T T
Ėgl = U · R · D · R · U ≈ R · D · R (7.39)
Next, the time derivative of the Cauchy stress in Equation (7.32) must be expressed as a func-
tion of the invariant stress rate to incorporate the constitutive equations in the formulation:
T T T
σ̇ = Ṙ · τ · R + R · τ̇ · R + R · τ · Ṙ (7.40)
T
The rotation tensor R is orthogonal (R = R−1 ), hence:
T T T
σ̇ = Ṙ · (R−1 · R) · τ · R + R · τ̇ · R + R · τ · (RT · R) · Ṙ
T T
= Ṙ · R−1 · σ + R · τ̇ · R + σ · R · Ṙ (7.41)
or:
T T T
R · τ̇ · R = σ̇ − Ṙ · R · σ − σ · R · Ṙ (7.42)
The above rate formulation is known as the Green-Naghdi rate. The Green-Naghdi rate equals
0 in case of rigid rotation (τ̇ = 0). In fact, the Green-Naghdi rate is a material derivative in
a co-rotating reference configuration.
Summarizing, the derivative of the Cauchy stress tensor can be expressed as a function of
the Cauchy stress tensor and the derivative of the invariant stress tensor τ , which, on its
turn is defined by the constitutive Equation (7.37). Now, it is possible to determine the time
derivate of the internal force vector, Equation (7.32), and similar to Equation (7.26), the
stiffness matrix for geometrically nonlinear behavior can be determined. (The derivation of
the stiffness matrix is beyond the scope of this course).
were introduced for the nonlinear finite element method where the concept of (pseudo-)time
was already introduced.
Therefore during an iteration of the Newton-Raphson scheme a trial solution for the displace-
ment increment is found. Based on the displacement increments strain increments at each
integration point in the Finite Element model is found using the strain-displacement matrix
of each element. The reason that only the values at the integration points are required is
due to the fact that the internal forces and the tangent stiffness matrix are computed by
taking the integrals given in equations (7.22) and (7.26), respectively, for which numerical
integration is used.
Within a time increment therefore the stress increment that corresponds to the given strain
increment must be computed that also satisfies all the constitutive relations that are for-
mulated for the chosen material. This procedure is called ‘stress update’. After computing
the stresses at the integration points directly the internal force vector can be computed and
convergence can be checked. If it is not attained then a new trial solution for displacement
must be found using the tangent stiffness matrix. To be able to do that however the relation
between the stress and strain increment is also required as shown in equations (7.26) and
(7.24).
In order to formulate a stress update algorithm first the discrete form of the constitutive
relations are presented. For elastoplastic material behavior, the elastic strain tensor is related
to the Cauchy stress tensor via the elasticity tensor E. When the total strain is decomposed
into a plastic part and an elastic part, the elastic stress-strain relation in incremental form
yields:
∆σ = E : ∆εe = E : (∆ε − ∆εp ) (7.43)
The plastic strain rate is integrated by a generalized trapezoidal rule, yielding the following
expression for the incremental plastic strain:
t(n+1)
Z
∆εp = ε̇p dt ≈ (1 − α)ε̇p(n) ∆t + αε̇p(n+1) ∆t (7.44)
t(n)
where t(n) and t(n+1) bound a time increment. The direction of the incremental plastic strain
is determined by interpolating the plastic strain rate directions at the start and end of the
time increment with the interpolation value α. Several well known integration techniques are
related to a fixed value of the interpolation value α. Taking α = 0 will lead to the Euler
forward method which is based on values at the start of the step only. This algorithm is
an explicit algorithm since the strain or stress of the current step has no influence on the
integration algorithm. Taking α = 1 will lead to the Euler backward method which is based
only on values at the end of the step and thus is a fully implicit method. In the remainder of
this section, the focus will be on the Euler Backward method since this integration scheme is
unconditionally stable, contrary to the Euler forward method, see Figure 7.4.
An expression for the plastic strain increment is obtained by substituting the definition of the
∂φ
flow rule (i.e. ε̇p = λ̇ ) into Equation (7.44) and taking α = 1:
∂σ
∂φ ∂φ
∆εp = ε̇p(n+1) ∆t = λ̇(n+1) ∆t ≈ ∆λ(n+1) (7.45)
∂σ(n+1) ∂σ (n+1)
7.6 Stress update 93
Then, substitution of Equation (7.45) in (7.43) yields the expression for the current stress
state:
σ (n+1) = σ(n) + ∆σ
∂φ
= σ(n) + E : ∆ε − ∆λ(n+1) E : (7.46)
∂σ (n+1)
The new stress state σ (n+1) is determined using an elastic predictor / plastic corrector method.
The basic idea is to treat the total strain increment as elastic, calculate the corresponding
stress, and then project this calculated stress onto the yield surface. The elastic predictor
defines a trial stress state σ t :
σt = σ (n) + E : ∆ε (7.47)
σ (n) is the stress state at time t(n) , see Figure 7.4. If the trial stress lies within the yield
surface, then φ(σ t , εpeq(n) ) ≤ 0, and the trial stress will be the new stress state: σ (n+1) = σ t .
If the trial stress state lies outside the yield surface then φ(σ t , εpeq(n) ) > 0. Since the yield
function can never be larger than 0, a plastic corrector has to be used to determine a new
stress state σ (n+1) which lies on the yield surface φ(σ (n+1) , εpeq(n+1) ) = 0. In case of hardening,
the yield surface changed due to plastic deformation, see dashed line in Figure 7.4.
Note: This figure also shows why the Euler forward method is conditionally stable. The
Euler forward method is based on values at the start of the increment, so the direction in
which the trial stress is scaled to the yield surface equals ∂σ∂φ . Therefore it is possible that
(n)
the return mapping algorithm cannot find a solution for the new stress state σ̃ (n+1) , using
the Euler forward stress update.
The equations presented above can be solved using the conventional backward-Euler approach
by defining a set of non-linear equations to be satisfied at time (n+1). However this requires a
non-linear solution scheme such as the Newton-Raphson method for which second derivatives
of the yield function with respect to stress and strain are required. However other algorithms
also exist that circumvent this thereby reducing the complexity of implementation. In the
following one of these algorithms is presented.
94 Finite Element Method
It is clear that the plastic multiplier ∆λ must be known to determine the new stress state.
The value of ∆λ can be determined by a Taylor expansion of the yield function around the
trial stress state σ t :
∂φ ∂φ
φ(σ (n+1) , εpeq(n+1) ) = φ(σ t , εpeq(n) ) + : ∆σ + p ∆εp + O(∆2 ) = 0 (7.50)
∂σ t ∂εeq(n) eq(n)
Neglecting the higher order terms (O(∆2 )) and taking into account that ∆σ = σ (n+1) − σt
gives an expression for the plastic multiplier ∆λ:
∂φ ∂φ ∂φ
φ(σ t , εpeq(n) ) − ∆λ :E: + p ∆εpeq(n) = 0 (7.51)
∂σ t ∂σ t ∂εeq(n)
The equivalent plastic strain can be written as a function of the plastic multiplier. Starting
point is that the incremental plastic deformation energy can be defined by the product of
the incremental strain tensor and the stress tensor, as well as by the incremental equivalent
plastic strain multiplied by the equivalent stress:
∂φ ∆λ ∂φ
σy ∆εpeq = σ : ∆εp = ∆λσ : ⇔ ∆εpeq = σ: (7.52)
∂σ σy ∂σ
Substitution of Equation (7.52) into (7.51) gives an expression for the plastic multiplier. The
plastic multiplier can be calculated and subsequently, the new stress state σ (n+1) is obtained
by Equation (7.48).
Due to the inaccuracy introduced by the linearization given in (7.50) however it is expected
that the yield function will be violated after one iteration. However, replacing the trial stress
with the newly found stress and repeating the process will give an iterative approach in
searching for the solution without requiring second derivatives. Below an example is given
for the cutting-plane algorithm applied to Von Mises plasticity.
Substitution of Equation (7.54) into Equation (7.52) yields a relation between the plastic
multiplier ∆λ and the equivalent plastic strain increment ∆εpeq :
√
∆εpeq = 2∆λ (7.55)
Substitution of Equation (7.55) into Equation (7.51) and taking into account the chain rule
∂φ ∂σy
for ∂φ
∂ε = ∂σy ∂ε gives:
∂φ ∂φ ∂φ ∂σy √
φ(σ (k−1) , εp(k−1)
eq ) − ∆λ(k) : E : + 2∆λ(k) = 0 (7.56)
∂σ (k−1) ∂σ (k−1) ∂σy(k−1) ∂εp(k−1)
eq
Note that the trial stress is replaced by a stress state with superscript (k−1) . This is because
the above expression will be used in an iterative procedure where (k) denotes the current
iteration. The stress σ(k−1) for k = 1 equals the trial stress σ t . Rewriting the above equation
gives an expression for the plastic multiplier in iteration k:
p(k−1)
φ(σ (k−1) , εeq )
∆λ(k) = (7.57)
∂φ ∂φ ∂σy
:E: + 2 p(k−1)
∂σ (k−1) ∂σ (k−1)
∂εeq
A new stress state in the kth iteration is determined analogous to equation (7.48):
∂φ
σ (k) = σt − ∆λE : (7.59)
∂σ (k−1)
Note: the direction of the stress projection (direction of the plastic strain increment) depends
on the stress state of the previous iteration (σ (k−1) ). In the limit, the direction of the
plastic strain increment depending on the new stress state (σ (k) ) and the one depending
on the stress state of the previous iteration will be almost equal. This means that the new
stress state depends on the direction of plastic strain increment of the new stress state,
which is characteristic for an Euler Backward method. Also note that the new stress state
is is determined by projecting the stress state from the trial stress, using the total plastic
multiplier. If the new stress state σ (k) fulfills the yield criterion, then the new stress state
96 Finite Element Method
σ (n+1) = σ (k) and the iterative process can be stopped. Summarizing, the iterative stress
update algorithm looks as follows:
initialize
σy = σy(n−1)
dσy
determine
dεpeq
k=0
σ (k) = σ t
determine φ(σ (k) , σy )
p(k)
WHILE (kφ σ (k) , σy (εeq ) k > tolerance) DO
k := k + 1
determine ∆λ(k)
∆λ(k)
P
determine ∆λ =
p(k) √
determine ∆εeq = 2∆λ
p(k) p(k)
determine εeq = εpeq(n) + ∆εeq
p(k)
determine σy (εeq )
dσy
determine p(k)
dεeq
∂φ
determine σ (k) = σ t − ∆λE :
∂σ (k−1)
END
Note that the subscript (n) is omitted to enhance readability.
Note that the initial residual force vector is the external force vector itself, because the
internal force vector is {0} in the first iteration (k = 1). Additionally, the deviation between
the internal force vector and external force vector of time step (n) can be accounted for in
time step (n+1), by introducing this difference as an additional external force. This process is
y(k)
called load correction. The material matrix [L(n) ] represents elastic or elastic–plastic material
behavior. The B-tensor is here formulated as a matrix, contrary to the formulation given in
Equation (7.14) where it is defined as a 3rd -order tensor. In the intermezzo below, it is shown
how the [B]-matrix can be constructed.
As an illustration, the [B]-matrix will be derived for a bi-linear rectangular element The
element has a length of 2a in x-direction and a length of 2b in y-direction. The origin of
the coordinate system is located in the middle of the element. In matrix-vector notation,
Equation (7.10) can be written as:
{v} = [N ]{v̄}
vx1
vy1
( ) " #
vx N1 0 N2 0 N3 0 N4 0
= vx2 (7.63)
vy 0 N1 0 N2 0 N3 0 N4
..
.
vy4
(a − x)(b − y) (a + x)(b − y)
N1 = N2 =
4ab 4ab
(a + x)(b + y) (a − x)(b + y)
N3 = N4 = (7.64)
4ab 4ab
Note that the vector with nodal velocities and the interpolation functions are defined differ-
ently compared to Section 7.3.
The rate of deformation vector reads:
∂
0
Dx ∂x
( )
v
∂ x
Dy = 0 (7.65)
∂y
∂ ∂ vy
D 1 1
xy 2 ∂y 2 ∂x
Substitution of Equation (7.63) in (7.65) gives the expression for the [B]-matrix ({D} =
[B]{v̄}):
−(b − y) 0 (b − y) 0 (b + y) 0 −(b + y) 0
1
[B] = 0 −(a − x) 0 −(a + x) 0 (a + x) 0 (a − x)
4ab
−(a − x) −(b − y) −(a + x) (b − y) (a + x) (b + y) (a − x) −(b + y)
(7.66)
(k+1)
Solving Equation (7.60) gives an expression for {∆∆u(n) }.
98 Finite Element Method
(k+1) (k+1)
{x(n) } = {x(n−1) } + {∆u(n) } (7.68)
(k)
The incremental displacement {∆u(n) } is {0} in the first iteration (k = 1).
Then, the incremental strain is obtained via:
(k+1) (k+1) (k+1) (k+1) (k+1) (k+1)
{∆ε(n) } = {D(n) }∆t = [B(n) ]{v(n) }∆t = [B(n) ]{∆u(n) } (7.69)
and subsequently, the accompanying stress is calculated by a stress update algorithm (see
Section 7.6):
(k+1) (k+1)
{σ(n) } = f {∆ε(n) }, {h} (7.70)
with {h} a vector with history variables. Hence, the internal reaction force is obtained
through: Z
(k+1) (k+1) T (k+1)
{F(n) }int = [B(n) ] {σ(n) } dV (7.71)
V
(k+1) (k+1)
If the norm of the residual force vector {R(n) } = {F(n) }ext −{F(n) }int is larger than a user
specified tolerance, the Newton–Raphson iterative procedure will be continued: k := k + 1.
The Newton–Raphson loop is continued until the norm of the residual force vector is smaller
than the user specified tolerance. Subsequently, the data has to be prepared for the next
time increment, i.e. the results at the end of increment (n) will now be used as the starting
conditions for increment (n + 1).
Chapter 8
8.1 Introduction
In previous chapters, we considered elastic and (visco)plastic deformations. According to the
presented elastic or plastic material models, infinitely large deformations can be imposed.
A real material will fail earlier by fracture. Before final fracture the material locally looses
cohesion and microscopic voids can be detected. Upon further loading these voids grow in
size and neighbouring voids coalesce and form micro-cracks. These voids and micro-cracks
are also known as damage.
In Figure 8.1 a micrograph is presented of a cross section through the necked region of a tensile
specimen from a dual phase steel. The black dots are voids. You can see that there is not
Figure 8.1: Voids (black) in a DP600 steel, close to a tensile fracture on the right-hand side
of the picture.
99
100 Continuum Damage Mechanics
one void where the final fracture initiates, but the voids are distributed over a large region.
The models that are discussed in this chapter consider these distributed voids as part of
the continuum and are therefore classified as Continuum damage mechanics (CDM) models.
CDM models distinguish commonly between 3 stages: initiation, growth and coalescence of
voids. After coalescence, micro-cracks exist that grow into macroscopic cracks that lead to
fracture. Macroscopic cracks are typical discrete discontinuities in a material that cannot be
modelled with continuum models.
such that a critical value Dc is reached when the equivalent plastic strain reaches the fracture
strain:
D(εpf ) = Dc (8.4)
Some famous fracture criteria use the following functions in (8.3):
(based on highest principal stress) Cockroft and Latham (1968) f (σ) = σ1 (8.5)
Rice and Tracey (1969) f (σ) = exp(Cη) (8.6)
Oyane (1980) f (σ) = 1 + Cη (8.7)
p B
Goijaerts (1999) f (σ) = h1 + Aηi (ε ) (8.8)
In the Cockroft and Latham criterion, the highest principal stress is used and integrated over
the equivalent strain. The Rice and Tracey model directly use the triaxiality and gives it
an exponential weight in the integration. The Oyane model assumes a linear influence of
8.3 Coupled damage models 101
the triaxiality. Goijaerts adapted this model by using the Macaulay brackets (hxi equals x
if x > 0 and equals 0 if x ≤ 0). This assumes that no damage evolution takes place under
negative triaxiality. Furthermore, a better fit could be obtained by weighing with a power
function of the equivalent plastic strain.
Because the uncoupled damage models are not directly taking the damage into account while
plastic deformation takes place, these models are especially suited for ductile materials, where
significant plastic deformation takes place before damage evolves.
Using this model, the material will have the full, undamaged, stiffness if the damage variable
has the value zero and it will have no stiffness left if the damage variable has the value one.
It then has no remaining load bearing capacity.
It can be expected that voids in a material will affect both the stress and strain, compared
to an unvoided material. For modelling purposes, simplifying assumptions are made. Two
different strategies are known as the stress equivalence principle and the strain equivalence
principle. According to the stress equivalence principle, the nominal stress and the effective
stress are equal and according to the strain equivalence principle, the nominal strain and the
effective strain are equal. The strain equivalence principle is mostly used and also adopted
here, hence:
ε = εeff (This is a second order tensor) (8.10)
In the undamaged material, we assume the Hookean relation σ eff = Eeff : εeff , where Eeff is
proportional with the undamaged Young’s modulus Eeff . For a damaged material we then
get
σ = E : ε = (1 − D)Eeff : εeff = (1 − D)σ eff (8.11)
An alternative interpretation is that the nominal stress and effective stress are related by the
effective load bearing area. In a uniaxial tensile test, the effective stress is the load divided
by the effective load bearing area, while the nominal stress is the load divided by the nominal
area. The damage variable can be considered as the void volume fraction.
(Void volume fraction)
Vd
Veff = V − Vd = 1− V = (1 − D)V (8.12)
V
102 Continuum Damage Mechanics
σf
σ Eeff
ε0 εmax
ε
where Vd is the void volume and V is the nominal volume. For spherical voids the void volume
fraction equals the void area fraction in a cross section:
such that
Aeff
F = σA = (1 − D)σeff = σeff Aeff (8.14)
1−D
Damage, however, can only grow and if the material is unloaded, the damage remains constant.
Therefore a parameter κ is introduced that represents the maximum obtained non-elastic
strain.
κ(t) = max(ε(τ )) 0 ≤ τ ≤ t (8.16)
In Equation (8.15), ε is then substituted by κ
The nominal Young’s modulus is derived from Equation (8.9). If the material is unloaded
from a damaged state, the stress–strain relation follows the line of reduced stiffness E, down
to the origin. This behaviour is depicted in Figure 8.2.
8.3 Coupled damage models 103
As long as the stress has not reached the fracture stress σf , the material is undamaged. When
the fracture stress is reached, micro-cracks initiate and evolve, leading to a reduced load
bearing capacity. Only when the strain reaches a value of εmax the load bearing capacity is
completely lost.
In the decreasing part of the stress–strain curve, the stress is derived from
κ − ε0
σ = (1 − D)Eeff κ = 1 − Eeff κ (8.17)
εmax − ε0
It should be noted that a negative slope in the stress–strain relation will result in mesh size
dependent results in finite element simulations and it requires a regularization or the use of
mesh size dependent ‘material’ parameters to get physically relevant results.
The approach for a uniaxial brittle damage model can easily be extended to the general case
by making the damage evolution dependent on an equivalent strain measure. It is however
questionable if damage due to micro-cracks should be modelled with isotropic damage or with
a more complex anisotropic damage model. In the latter case the damage variable D should
be written as a second, or even a fourth order tensor. This is beyond the scope of this course.
σf is the current flow stress, fc and ff are the void volume fractions at start of coalescence
of voids and at final rupture respectively. The model parameters q1 and q2 were added for a
104 Continuum Damage Mechanics
better fit. Setting them equal to one and using f = f ∗ degenerates the GTN model to the
original Gurson model.
The void volume evolution follows from the change of volume of the material, because the
plastic deformation of the matrix material is considered incompressible. This results in
˙ 2 3p
f = λ̇(f − f )σf sinh (8.21)
2σf
This term describes the growth of existing voids. Often an additional term for void initiation
is used.
Contrary to the original damage model by Kachanov, in the case of ductile damage, the
Young’s modulus is often taken constant and is not affected by damage.
Chapter 9
In this chapter the combined effect of material hardening and geometrical changes on the
stability of deformation is considered for planar stretching of sheet metal. The stability
analysis resembles the analysis for a tensile bar, but it is now extended to stretching of sheet
under bi-axial stress states. The analysis is relevant for determination of the formability limits
in sheet metal.
The limiting values for the deformation of sheet metal depend on many parameters. One
important type of failure is the fracture of sheet during the forming process. This fracture is
preceded by localized necking and depends on the current state of deformation. In the plane
of the sheet, the deformation in two directions influence the failure behavior, e.g. εx = 0.2
and εy = 0.2 is a completely different situation than εx = 0.2 and εy = −0.2. The first is a
state of equi-biaxial stretch and the second a state of pure shear.
In a uniaxial tensile test localized necking is preceded by diffuse necking. After a small
introduction to plasticity, this type of necking is treated first. Thereafter localized necking
for different proportional strain paths is described.
εx + εy + εz = 0 (9.1)
In a tensile test, we observe that after initial yielding of material, the tensile stress increases.
This is called work hardening and is often modelled with the Hollomon law:
where σy is the flow stress at a particular strain ε. For many steels this is a good approximation
for a large part of the hardening curve. To model the initial yield stress, an initial strain ε0
105
106 Tensile instability in sheet metal
Figure 9.1: Sheet forming instabilities: (a) uniform deformation, (b) diffuse necking and (c)
local necking.
Plastic deformation takes place if the equivalent stress reaches the value of the flow stress.
The strain ε in the Hollomon equation is then substituted by an equivalent strain that is
derived from the principle of energy equivalence:
Z t
εeq (t) = ε̇eq (τ ) dτ and σeq ε̇eq = σij ε̇ij (9.5)
0
In practice the Von Mises yield locus appears to be too rounded. To describe the higher cur-
vature that is often found near the uniaxial and equi-biaxial stress states, Hosford developed
an isotropic non-quadratic model in 1972 [4]. For the plane stress case the model can be
written as:
σeq = 21 (|σ1 − σ2 |m + |σ1 |m + |σ2 |m )1/m (9.6)
Experience shows good results for b.c.c. metals with m = 6 and for f.c.c. metals with m = 8.
The plane stress yield loci for m = 2 (Von Mises) and for m = 6 and m = 8 are presented in
9.2.
9.2 Diffuse necking 107
σ2
σ1
m=2
m=6
m=8
Finally, the direction of plastic strain is perpendicular to the yield surface, or in other words
to the (hyper)plane of constant equivalent stress:
∂σeq
ε̇ij = λ (9.7)
∂σij
A0 l0 = Al with A = wt (9.8)
where the 1-direction indicates the tensile direction. The tensile load P on the specimen is
P = σ1 A (9.10)
or in differential form
dP = dσ1 A + σ1 dA (9.11)
At the load maximum we have dP = 0 and we obtain
dσ1 dA
=− = dε1 (9.12)
σ1 A
108 Tensile instability in sheet metal
or
dσ1
= σ1 (9.13)
dε1
Applying the Hollomon hardening function σ1 = Cεn1 the condition for diffuse necking be-
comes:
nCε1n−1 = Cεn1 ⇒ n = ε∗1 (9.14)
The value ε∗1 is called the uniform strain. For many steels the value of n is in the order of 0.2
and the preceding formula indicates that a uniaxial tensile specimen will yield in a uniform
way up to a natural strain of about 0.2. At that point diffuse necking will start as indicated
in 9.1.b. In the load–deflection curve this point coincides with the maximum load.
Figure 9.3: FLD determination using the strip technique of Nakazima et al [8].
stress and principal strain directions to be equal restricts the validity of the model to planar
isotropic material behavior.
σ2 = ασ1 , σ3 = 0 (9.15)
dε2 = β dε1 , dε3 = −(1 + β) dε1 (9.16)
The unit forces (per sheet length) are defined as
Local necking is assumed to start if dT = 0. Due to the proportional loading, this will occur
for T1 and T2 simultaneously. We now have
dT1 = dtσ1 + t dσ1 = t(σ1 dε3 + dσ1 ) = t[−σ1 (1 + β) dε1 + dσ1 ] = 0 (9.18)
dσ1
= (1 + β)σ1 (9.19)
dε1
If the hardening of the material can be described by the Hollomon relation (9.2) it can be
derived that for a proportional deformation process the relation
σ1 = C ′ εn1 (9.20)
110 Tensile instability in sheet metal
holds, where C ′ = Cf (β, n). In this case the localized necking condition (9.19) becomes
nσ1
nC ′ ε1n−1 = = (1 + β)σ1 (9.21)
ε1
In a forming limit diagram this condition can be represented by a straight line through the
point where the major strain is n and the minor strain is 0, parallel to the line ε1 = −ε2 .
For the Hollomon hardening function, the limit strains are independent of the yield function.
This is not the case for other hardening functions.
✻
σ2
θ t n 2 σ1
✛ ❑✯ ✻
✲1 ✲
If necking takes place in the described way, a localization band emerges at a specific angle to
the major strain direction (9.4). The straining in the band accelerates, while the strain outside
the band remains constant (neglecting elastic springback). In order to keep the displacements
inside and outside the band compatible, the strain in the band, in the direction along the band
must be zero. Therefore, since we have proportional loading up to the localization strain, the
band will develop in a direction where the strain is zero. The strain εtt in a direction with
angle θ to the major strain direction can be derived from the following equation (graphically
represented by the Mohr circle):
ε1 + ε2 1+β
cos 2θ = − =− (9.25)
ε1 − ε2 1−β
9.3 Localized necking 111
A
1 dεA
dεB
B 2 P
dεA dεB
Q dε2
A arctan α0
dε1A dε1B
φ=0
(a) Plate with imperfection. (b) Strain increments inside and outside the groove.
It can be concluded from this result that at values β > 0 no solution for θ exists. Therefore,
the localization criterion according to Hill is only valid for the extension–compression region
(the left side) of the FLD.
For uniaxial tensile stresses and isotropic material, the width and thickness strain are equal
and from the constant volume constraint we obtain
and therefore
dε2 1
β= =− (9.27)
dε1 2
For uniaxial loading of an isotropic material we will get a localization band at an angle
1+β 1 1
arccos − 31 ≈ 55◦
cos 2θ = − =− ⇒ θ= 2 (9.28)
1−β 3
Compatibility between the uniform part A and the groove B requires that
The force per unit sheet length in direction 1 must be transmitted through the groove, hence
Where f = tB /tA is the current thickness ratio. As long as f ≈ 1, the stress ratio α0
approximately holds for both regions A and B. Since the stress σ1B in the groove is larger
than σ1A in the uniform part, the material in the groove reaches the yield surface first. In
9.5(b) this is approximately at position P . Because of the constraint equation (9.31) no
yielding takes place, since the uniform region is still fully elastic. The stress state in region
B will move along the yield locus to position Q, until also region A reaches the yield locus
at position P . This situation is depicted in 9.5(b). In that period σ1B has increased and σ2B
has decreased, hence the stress ratio α has decreased and proportional deformation is not
possible in the groove.
With the stress state in A and B at two different positions on the yield locus, the normal to
the yield surface is different and because of the constraint (9.31), the strain perpendicular to
the neck must be larger in the neck than in the uniform part. This is presented in the vector
plot in 9.5(b). As a consequence, the hardening in region B will be larger and the two stress
situations can no longer be projected on the same yield locus. With an equal strain increment
in the 2-direction and a larger strain in the 1-direction, the thickness strain in the neck will
be more negative than in the uniform part, decreasing the thickness ratio (f < f0 ).
The analysis of the deformation can be further developed numerically. The drawing region
can be included in the analysis by assuming an inclined groove. For planar isotropic material
the inclination angle can be determined beforehand by (9.25). Strain increments ∆εA are
prescribed on region A, respecting the proportionality ratio β0 . For every increment ∆εA a
strain increment ∆εB is calculated iteratively, such that the plate forces normal and tangential
to the groove in region A and B are in equilibrium. The sheet is considered to have failed if
the strain increment in the groove is larger than a prescribed multiple of the strain increment
in the uniform part. The strain in the uniform region at the time of failure is considered to
be the limiting strain for that particular strain ratio β0 .
In 9.6 the strain paths in the plate (region A) and in the groove (region B) are presented in
a forming limit diagram. An initial thickness ratio f0 = 0.99 was used. The failure criterion
stopped the increments when the strain increment in the groove was more than 10 times
the strain increment in the uniform region. A Von Mises yield locus was used with a Swift
hardening model with C = 446.3 MPa, n = 0.292 and ε0 = 0.0039.
In the drawing region of the FLD, the strains inside the groove in the directions 1 and 2
have the same ratio as in the uniform region. The limit strain follows a line at an angle of
45◦ with respect to the major strain axis, as predicted by the Hill analysis. The limit strain
for the plane strain path is lower than the corresponding n-value of 0.292 because of the
9.3 Localized necking 113
0.7
0.6
0.5
major strain
0.4
0.3
0.2
0.1
plate
groove
FLC
0
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6
minor strain
Figure 9.6: The strain-paths inside and outside the groove in a Marciniak–Kuczynski analysis.
Figure 9.7: The Hill theory and the Marciniak–Kuczynski theory against experiment (◦) [8].
initial thickness ratio of 0.99. In the stretching region, the strains inside and outside the
groove are no longer in the same direction. It can be seen that the strain ratio ε1B /ε1A keeps
increasing until the rate of deformation in the groove approaches a plane strain situation. For
a fixed strain increment in part A, the strain increment in the groove becomes very large and
localisation is detected.
The results of the Marciniak–Kuczynski analysis are very sensitive to the initial thickness
ratio f0 (see 9.7). If experimental FLC values are known for the plane strain condition, this
is often used to calibrate f0 . The estimation at the equi-biaxial point is shown to be very
sensitive to the used yield function [9]. If a Von Mises yield function is used, the FLCs tend to
be much too high near the equi-biaxial direction (see 9.7). It can be appreciated from 9.5(b)
114 Tensile instability in sheet metal
0.7
0.6
0.5
major strain
0.4
0.3
0.2
m=2
0.1
m=6
m=8
0
-0.3 -0.2 -0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
minor strain
Figure 9.8: Predicted FLCs using the Hosford yield function and f0 = 0.995.
that the local curvature of the yield locus is very important. Indeed, if the Hosford yield
function is used for representative values of m the shape of the predicted FLC becomes much
better, as can be seen in 9.8. Notice that—at least for the Swift hardening model—the left-
hand side of the FLC is not influenced by the shape of the yield function. This corresponds
to the results of the Hill local necking analysis.
-- END OF THEORY --
Appendix A
In each point, at least three mutual perpendicular planes can be defined on which only a
normal stress acts. Then, the stress vector T has the direction of the normal vector:
T=σ·n (A.1)
The size of the principal stress is σ. The components of the stress vector can be written as
(see equation 4.3):
Tx = σ cos (α) = σxx cos (α) + σxy cos (β) + σxz cos (γ)
Ty = σ cos (β) = σxy cos (α) + σyy cos (β) + σyz cos (γ) (A.2)
Tz = σ cos (γ) = σxz cos (α) + σyz cos (β) + σzz cos (γ)
or in eigen-value representation:
σxx − σ σxy σxz
cos (α)
0
σxy
σ yy − σ σ yz
·
cos (β) = 0 (A.3)
σxz σyz σzz − σ cos (γ) 0
⇐⇒ |σij − σδij | = 0
Working out this equation yields a secular equation of the third degree, with three roots. The
three roots represent the principal stresses. By definition it is stated that: σ1 ≥ σ2 ≥ σ3 .
Elaboration of the determinant gives:
σ 3 − I1 σ 2 − I2 σ − I3 = 0 (A.5)
115
116 Principal stresses and invariants
The coefficients of this equation are the invariants of the stress tensor:
The sign ”:” means double contraction, comparable to the inner product of 2 vectors. In
index notation, A : B is written as Aij Bij
Deviatoric stress
The first invariant, the sum of the the principal stresses, equals the isotropic part of the
stress tensor:
1 1 1
3 I1 = 3 tr (σ) = 3 (σ1 + σ2 + σ3 ) = −p (A.7)
with p the hydrostatic pressure (average pressure). Next, σ can be split into an isotropic part
and a part of which the average normal stress is zero:
s = σ − 13 tr (σ)1 = 4 I − 13 11 : σ
(A.9)
Three equations form the foundation of the graphical representation of Mohr’s circles. If
the coordinate system is placed on the principal axis, then equation (A.2) degenerates to:
With σn = |s · n| = |sx cos (α) + sy cos (β) + sz cos (γ)| one can write:
an expression for the shear stress components can be found by using the theorem of Pythagoras
and the symmetry condition of the stress tensor σij = σji :
τ 2 = |τ |2 = |s|2 − σn2
(A.14)
τ 2 = σ12 cos2 (α) + σ22 cos2 (β) + σ32 cos2 (γ) − σn2
From this system of equations, 2 unknowns can be eliminated: 1. Elimination of cos(β) and
cos(γ) gives:
(σn − A1 )2 + τ 2 = R12 (A.16)
This equation represents a circle with:
A1 = 12 (σ1 + σ2 ) (center)
q
R1 = (σ1 − σ2 )(σ1 − σ3 ) cos2 (α) + 41 (σ1 − σ2 )2 (radius, min. for cos(α) = 0)
with:
A2 = 21 (σ1 + σ2 ) (center)
q
R2 = (σ3 − σ1 )(σ3 − σ2 ) cos2 (γ) + 41 (σ1 − σ2 )2 (radius, min. for cos(γ) = 0)
with:
A3 = 21 (σ1 + σ3 ) (center)
q
R3 = (σ2 − σ1 )(σ2 − σ3 ) cos2 (β) + 14 (σ1 − σ3 )2 (radius, min. for cos(β) = 0)
Three extremes are found for cos(α), cos(β) and cos(γ) = 0, respectively, i.e. Mohr’s stress
circles. The graphical representation of Mohr’s stress circles is depicted in figure A.1. The
stresses acting on an arbitrary plane are somewhere in the hatched areas.
The determination of the principal strain is analogue to the determination of the principal
stresses:
εxx − ε εxy εxz
det
εxy εyy − ε εyz =0 (A.19)
εxz εyz εzz − ε
Elaboration of these equations ends up with a secular equation with three roots: i.e. the
principal strains ε1 ≥ ε2 ≥ ε3 . By definition it is stated that: ε1 ≥ ε2 ≥ ε3 . Elaboration of
the determinant gives:
ε3 − J1 ε2 − J2 ε − J3 = 0 (A.20)
The invariants of the strain tensor read:
J1 = tr (ε) = εxx + εyy + εzz = ε1 + ε2 + ε3
J2 = 12 (ε : ε − (tr (ε))2 ) = −εxx εyy + εyy εzz + εzz εxx + (ε2xy + ε2yz + ε2zx )
(A.21)
= −ε1 ε2 + ε2 ε3 + ε3 ε1
J3 = det(ε) = ε1 ε2 ε3
Deviatoric strain
The first invariant, the sum of the principal strains, equals the isotropic part of the strain
tensor. This invariant is related to the specific volume change for small strains:
∆V V − V0
= = (1 + εxx )(1 + εyy )(1 + εzz ) − 1 ≈ εxx + εyy + εzz = J1 (A.22)
V0 V0
Next, ε can be split into a part which involves volume changes (the isotropic part) and a part
that involves shape changes. The isotropic part of the latter is zero.
ε = e + 31 tr (ε)1 (A.23)
e is known as the deviatoric strain tensor and has the following invariants:
′
J1 = tr (e) = exx + eyy + ezz = e1 + e2 + e3
′
J2 = 21 (e : e − (tr (e))2 ) = −exx eyy + eyy ezz + ezz exx + (e2xy + e2yz + e2zx )
(A.24)
= −e1 e2 + e2 e3 + e3 e1
′
J3 = det(e) = e1 e2 e3
119
The determination of Mohr’s strain circles is analogue to the determination of Mohr’s stress
circles, and the result is depicted in figure A.2, where γ represents the shear strain:
[1] Bonet, J., and Wood, R. D. Nonlinear continuum mechanics for finite element
analysis, 2nd ed. Cambridge University Press, Cambridge, 2008.
[2] Hill, R. On discontinuous plastic states, with special reference to localized necking in
thin sheets. Journal of the Mechanics and Physics of Solids 1 (1952), 19–30.
[5] Kachanov, L. M. On the creep fracture time. (in russian). Izv. Acad. Nauk SSSR Otd.
Tekhn 8 (1958), 26–31.
[6] Krabiell, A., and Dahl, W. Zum Einfluss von Temperatur und Dehngeschwindigkeit
auf die Streckgrenze von Baustählen unterschiedlicher Festigkeit. Archiv für das
Eisenhüttenwesen 52 (1981), 429–436.
[7] Marciniak, Z., and Kuczynski, K. Limit strains in processing of stretch forming
sheet metal. Int. J. Mech. Sci. 9 (1967), 609–620.
[9] Vegter, H., An, Y., Pijlman, H. H., and Huétink, J. Different approaches to
describe the plastic material behaviour of steel and aluminium-alloys in sheet forming. In
Proceedings of the 2nd ESAFORM Conference on Material Forming (Guimarães, 1999),
J. A. Covas, Ed., pp. 127–132.
[10] Zener, C., and Hollomon, J. H. Effect of strain rate upon plastic flow of steel.
Journal of Applied Physics 15 (1944), 22.
121