高等固体力学
高等固体力学
高等固体力学
AND NANOMECHANICS
Lecture Notes (CE236/C214)
SHAOFAN LI
Department of Civil and Environmental Engineering,
University of California, Berkeley, CA94720, USA
University of California
Berkeley, California
Contents
1. INTRODUCTION 1
2. PRELIMINARY 4
2.1 Vectors and Tensors 4
2.2 Review of Linear Elasticity Theory 13
2.3 Exercises 19
3. HOMOGENIZATION I — CLASSICAL AVERAGING METHOD 21
3.1 Representative volume element 21
3.2 Average stress in an RVE 22
3.3 Average strain and strain rate 24
3.4 Definition of eigenstrain, eigenstress, and inclusion 26
3.5 Eshelby’s equivalent eigenstrain method I: Traction boundary
condition 27
3.6 Eshelby’s equivalent eigenstress method II: Displacement boundary
condition 28
3.7 Effective material properties via eigenstrain method 29
3.8 Jock Eshelby (I) 33
3.9 Exercises 35
4. GREEN’S FUNCTION AND FOURIER TRANSFORM 37
4.1 Green’s Function 37
4.2 Fourier transform 40
4.3 Examples of Green’s Function 46
4.4 Static Green’s function for 3D linear elasticity 49
4.5 Variation in a Theme: Radon Transform 54
4.6 Joseph Fourier(I) 58
vi INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
4.7 Exercises 61
5. EIGENSTRAIN THEORY 63
5.1 Fundamental equations of micro-elasticity 63
5.2 Method of Green’s Functions 66
5.3 Application I: Dislocation problems 68
5.4 Application II: Stress intensity factor for a flat ellipsoidal crack 71
5.5 Isotropic inclusion-Eshelby’s solution 78
5.6 Exterior Solution of Ellipsoidal Inclusion 83
5.7 Jock Eshelby (II): Lessons from J.D.Eshelby 88
5.8 Exercises 90
6. EFFECTIVE ELASTIC MODULUS 95
6.1 Effective elastic moduli for composites of dilute suspension 95
6.2 Self-consistent method 104
6.3 Mori-Tanaka methods 109
6.4 Rodney Hill 120
6.5 Exercises 124
7. INTRODUCTION OF DISLOCATION THEORY 127
7.1 Screw dislocation 127
7.2 Edge dislocation 134
7.3 The Peach-Koehle force 139
7.4 Configuration force: Eshelby’s energy-momentum tensor 144
7.5 Continuum theory of dislocation 150
7.6 Discrete Dislocation Dynamics (DD) 160
7.7 The Peierls-Nabarro Model 164
7.8 Dislocations in the epitaxial thin film 176
8. COMPARISON VARIATIONAL PRINCIPLES 188
8.1 Review of Variational Calculus 188
8.2 Extreme variational principles in linear elasticity 192
8.3 Hashin-Shtrikman variational principles 199
8.4 Review of Functional Analysis and Convex Analysis 205
8.5 Legendre Transformation and Duality 215
8.6 Legendre-Fenchel transformation in linear elasticity 222
8.7 Talbot-Willis variational principles 224
Contents vii
Chapter 1
INTRODUCTION
dia and even cooling of packed flowers (Van der Sman [1997][2000]); In later
1980s, Clementi and his co-workers [1988] initiated the idea of multiscale
modeling, or multiscale simulation, i.e. using super-computers to conduct
large scale computations that combine ab initio modeling, classical molecu-
lar dynamic modeling, and phenomenological modeling in a single simula-
tion. The unified macroscopic, atomistic, ab initio dynamics (MAAD) de-
scription brings all three descriptions together into a seamless union, embrac-
ing all the size scales, from the very small to the very big (e.g. Abraham et al
[1996],[1997ab],[2000]).
The simplest and earlist multi-scale modeling notion is the so-called Cauchy-
Born rule. By combining this concept with the finite element methods, the so-
called quasicontinuum method was developed by Tadmor, Ortiz, Phillips and
their co-workers (Tadmor et al 1996). The Cauchy-Born rule is ensentially a
simplistic “homogenization postulation” in lattice kinematics, and it serves as
passage to link between the molecular dynamics and continuum mechanics.
The Born rule assumes that the continuum energy density W can be computed
using an atomic potential, with the link to the continuum being the deformation
gradient F. To briefly review continuum mechanics, the deformation gradient
F maps an undeformed line segment dX in the reference configuration onto a
deformed line segment dx in the current configuration,
dx = FdX (1.1)
In general, F can be written as
du
F=I+ (1.2)
dX
where u is the displacement vector. If there is no displacement in the contin-
uum, the deformation gradient is equal to unity.
The major restriction and implication of the Cauchy-Born rule is that the
continuum deformation must be homogeneous. This results from the fact that
the underlying atomic system is forced to deform according to the contin-
uum deformation gradient F. By using the Born rule, one may be able to
derive a continuum stress tensor and tangent stiffness directly from the inter-
atomic potential, which allowed the usage of the standard nonlinear finite ele-
ment method. This procedure is now called as the so-called quasi-continuum
method.
Apparently, the contemporary mico-mechanics or nano-mechanics is only
at its infancy. There are many unknown approaches to be explored and many
new phenonmena to be studied. In this lecture notes, we are attempting to
synthesize the most recent research results in the forefront of nano-mechanics
while presenting traditional micro-mechanics in a coherent fashion. By doing
so, we hope that it may serve as a stepping stone for us to reach a new height
in the quest for a multiscale nano-mechanics of our time.
4 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Chapter 2
PRELIMINARY
x = x1 e1 + x2 e2 + x3 e3 = xi ei = (x · ei )ei (2.1)
where Einstein convention is used that the repeated indices indicates summa-
tion from 1 to 3.
Consider two vectors, V = Vi ei and W = Wj ej . The scalar (dot) product
of two vectore, V and W, is defined as
V · W = Vi ei · Wj ej = Vi Wj ei · ej = Vi Wj δij = Vi Wi (2.2)
where
1, i=j
ei · ej = =: δij (2.3)
0, i 6= j
is called Keronecker delta.
A cross product of two vectors, A = Ai ei , B = Bj ej , is defined as
This definition can be explained as a permutation rule that change of any two
adjcent indces of the symbol, there is a negative sign (−1) occurs.
For example, since e123 = 1, then
and
e312 = (−1)e132 = (−1)(−1)e123 = (−1)(−1)1 = 1
The cross product of two vectors can also written as
Since
e1 e2 e3
ei × ej = δ1i δ2i δ3i (2.8)
δ1j δ2j δ3j
then
δ1k δ2k δ3k δ1i δ2i δ3i
ekij = eijk = (ei × ej ) · ek = δ1i δ2i δ3i = δ1j δ2j δ3j (2.9)
δ1j δ2j δ3j δ1k δ2k δ3k
This provides a link between permutation symbol and Keronecker delta.
Consider the product of two permutation symbols,
δ1i δ2i δ3i δ1r δ2r δ3r
eijk erst = δ1j δ2j δ3j δ1s δ2s δ3s
δ1k δ2k δ3k δ1t δ2t δ3t
δ1i δ2i δ3i δ1r δ2s δ3t
= δ1j δ2j δ3j δ1r δ2s δ3t
δ1k δ2k δ3k δ1r δ2s δ3t
δir δis δit
= δjr δjs δjt (2.10)
δkr δks δkt
One may show that for any second order tensor A,
6 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
The dyad is called the second order tensor 1 , and its basis, ei ⊗ ej , is called
dyadic basis. In this case, the components of the second order tensor are Cij =
Ai Bj .
In fact, every second order tensor can be expressed in a dyadic basis, such
as
σ = σij ei ⊗ ej (2.12)
= ij ei ⊗ ej (2.13)
This is why in linear elasticity we may define the infinitesimal strain tensor as
1 1
= ∇ ⊗ u + (∇ ⊗ u)T = uj,i + ui,j ei ⊗ ej (2.15)
2 2
1
or in component form ij = uj,i + ui,j .
2
In general, a n-th order tensor is a polyads, or has a polyadic representation,
e.g.
C = Cijk` ei ⊗ ej ⊗ ek ⊗ e` (2.16)
is a forth order tensor.
Analogous to the scalar product of vectors, the double contraction of two
tensors are defined as two dot products among of Cartesian tensor bases, i.e. if
A = Aij ei ⊗ ej and B = Bk` ek ⊗ e` , then
In each contraction, there are two bases annihilated. Consider a forth order
tensor C = Cijk` ei ⊗ ej ⊗ ek ⊗ e` and a second order tensor = ij ei ⊗ ej .
There are total six basis vectors. A double contraction between the two will
annihilate four basis vectors and produce a second order tensor, i.e.
σ = C : = Cijk` ei ⊗ ej ⊗ ek ⊗ eell : st es ⊗ et
= Cijk` st ei ⊗ ej δks δ`t = Cijk` k` ei ⊗ ej (2.19)
eij = ei ⊗ ej . (2.23)
The second order unit tensor and the forth order unit tensor are constructed
based on the following rules:
1(2) := ei · ej ei ⊗ ej = δij ei ⊗ ej = δij eij (2.24)
1(4) := ei ⊗ ej : ek ⊗ e` ei ⊗ ej ⊗ ek ⊗ e`
= (eij : ek` )eij ⊗ ek` = δik δj` ei ⊗ ej ⊗ ek ⊗ e` (2.25)
The superscript indicates the order. It is interesting to note that the fourth order
unit tensor defined in (2.25) is not symmetric with all indices.
To represent symmetric tensors, it may be expedient to first define symmec-
tric tensor basis. The second order symmetric basis is defined as
1 1
eSij = eij + eTij = ei ⊗ ej + ej ⊗ ei (2.26)
2 2
Any second order symmetric tensor can then be expressed as S = Sij eSij . One
may denote the space of all second order symmetric tensors as
and
Q−1 = hE(1) + vE(2) (2.43)
By definition,
TΩ = 1(4s) − C : SΩ : D
= (E(1) + E(2) ) − (3KE(1) + 2µE(2) ) : (s1 E(1) + s2 E(2) )
1 1 (2)
: E(1) + E
3K 2µ
= (1 − s1 )E(1) + (1 − s2 )E(2)
The gradient of a vector field, a first order tensor field, is a second order tensor.
In general, the gradient operation increases the ordero f a tensorial field up to
one order higher.
On the other hand, the scalar product or contraction between a gradient op-
erator and a tensorial field is called divergence operation, which will result a
new tensorial field with reduced order. Consider a vector field, A = Ai ei . Its
divergence is being defined as
∂ ∂A ∂Ai
j
divA := ∇ · A = ei · Aj ej = (ei · ej ) = (2.49)
∂xi ∂xi ∂xi
The cross product between the gradient operator and a tensorial field. A =
Ai ei , is called the Curls or rot of the tensorial field.
∂Aj
CurlA := ∇ × A = ei × ej = eijk ∂i Aj ek = eijk ∂j Ak ek (2.50)
∂xi
In what follows, a few integral transformations are listed.
Suppose that there is a continuous function, f (x) ∈ C 1 (Ω), defined in a
domain Ω ∈ IRd with smooth boundary ∂Ω. A well-known integral theorem is
Z Z
∇f dΩ = f ndS (2.51)
Ω ∂Ω
or in component form
Z Z
∂f
dΩ = f ni dS (2.52)
Ω ∂xi ∂Ω
∂2u
∇ · σ + ρb = ρ (2.58)
∂t2
For convenience, we often write the component form
∂ 2 ui
σji,j + ρbi = (2.59)
∂t2
∂uji
where uji,j = .
∂xj
• Geometric relation
The infinitesimal strain field = ij ei ⊗ ej is defined as
1
= ∇ ⊗ u + (∇ ⊗ u)T (2.60)
2
Note that ∇ ⊗ u = uj,i ei ⊗ ej . Hence (∇ ⊗ u)T = ui,j ei ⊗ ej .
Therefore in component form,
1
ij = (ui,j + uj,i ) (2.61)
2
• Constitutive equations
For linear elastic solids, the constitutive equations have the following form,
Based on foundamental theorem of calculus, one may find its inverse relation-
ship as
∂U ∂U
= σ, = σij (2.71)
∂ ∂ij
The complementary strain energy density can be obtained via Legendre
transform,
U ∗ (σ) = σ : − U () (2.72)
Or one may define Z σ
0 0
U ∗ (σ) = (σ )dσ (2.73)
0
Preliminary 15
Proof:
Consider both states being equilibrium states. It has
Z Z
(1) (2) (1) (2)
fi ui dΩ = − σji,j ui dΩ
Ω ZΩ Z
(1) (2) (1) (2)
= − σji nj ui dS + σji ui,j dΩ
Z∂Ω Z Ω
(1) (2) (1) (2)
= − ti ui dS + σji ji dΩ (2.84)
∂Ω Ω
Moving the first term of the right-hand side of (2.74) to the left-hand side yields
Z Z Z
(1) (2) (1) (2) (1) (2)
fi ui dΩ + ti ui dS = σij ij dΩ (2.85)
Ω ∂Ω Ω
Consider the fact that the two systems exist in the same material
Z Z Z Z
(1) (2) (1) (2) (1) (2) (1) (2)
σij ij dΩ = Cijkl kl ij dΩ = Cklij kl ij dΩ = kl σkl dΩ
Ω Ω Ω Ω
Compare the both sides of (2.75) and (2.76), the theorem holds.
In addition, the equality
Z Z
(1) (2) (2) (1)
σij ij dΩ = σij ij dΩ (2.87)
Ω Ω
2.3 Exercises
Probelm 2.1 Let δu be a virtual displacement field and σ be a self-equilibrium
stress field. Show
∇ · σ · δu = ∇ · σ · δu − σ : (∇ ⊗ δu) (2.103)
Probelm 2.3 Suppose that there are two different solutions of equilibrium
equation,
∇ · σ 1 = 0, ∇ · σ 2 = 0 (2.106)
which satisfy the same boundary conditions,
u1 = u0 ,
∀x ∈ Γu (2.107)
u2 = u0 ;
n · σ 1 = t0 ,
∀x ∈ Γt (2.108)
n · σ 2 = t0 ;
S
where Γu Γt = ∂Ω.
20 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
∆σ = σ 1 − σ 2 = 0 (2.109)
∆ = 1 − 2 = 0 (2.110)
1(4) : A → A (2.111)
1
1(4s) : A → A + AT (2.112)
2
1
1(4a) : A → A − AT (2.113)
2
Homogenization I — Classical Averaging Method 21
Chapter 3
"Curiouser and curiouser!" cried Alice,"Now I’m opening out like the largest
telecope that everwas!"
— Lewis Carroll, Alice in Wonderland
∇ · σ = 0 ⇒ σij,j = 0. (3.1)
fined as Z
1
< T >X := T(x, X)dVx (3.2)
V V
If the material is homogeneous at macro-level, we have
Z
1
< T >:= T(x)dVx (3.3)
V V
For instance, if T = σ(x) is a micro-stress field, the macro-stress at a
material point will be Σ =< σ >. Similarly, if T = (x) is a micro-strain
field, the macro-strain at a material point is E =< >.
A very useful average theorem about micro-Cauchy stress tensor may be
stated as follows:
Theorem 3.1 Suppose an RVE is subjected to natural boundary condition,
and the traction on remote boundary of an RVE (∂V ) is generated by a constant
stress tensor, σ 0 . Then the average stress at this material point, or the macro
stress at the material point,
Σ =< σ >= σ 0 (3.4)
Note that the point here is that one only knows the traction distribution on the
remote boundary of the RVE, but one does not know the exact stress distribu-
tion inside the RVE.
Proof
Consider,
∂xi
= δij and σji,j = 0 (3.5)
∂xj
One then can express Cauchy stess inside an RVE as
∂x
j
σij = σik δkj = σik δjk = σik
∂xk
= (σik xj ),k − σik,k xj = (σik xj ),k (3.6)
Therefore,
Z Z
1 1
< σij > = σij dV = σik xj dV
V V V V ,k
I I
1 1
= σik xj nk dS = σ 0 xj nk dS
V ∂V V ∂V ik
0 I
σik 0 Z
σik ∂xj
= xj nk dS = dV
V ∂V V V ∂xk
0 Z
σik σ0 0
= δjk dV = ik δjk V = σij (3.7)
V V V
24 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
One can find the average displacement gradient field in terms of boundary data,
i.e., Z Z
1 1
< ui,j >= ui,j dV = nj u0i dS (3.9)
V V V ∂V
Note that you don’t know exact distribution of the displacement field inside the
RVE.
Moreover, one may find the average strain and rotation fields in terms of
boundary displacement data,
I
1 1
< ij >= < ui,j > + < uj,i > = (nj u0i + ni u0j )dS
2 2V ∂V
I
1 1
< ωij >= < ui,j > − < uj,i > = (nj u0i − ni u0j )dS
2 2V ∂V
Remark 3.3.1 in general, the average displacement fields of an RVE can
not be expressed in terms of remote surface data. To see this, one may evaluate
the average displacement field. Using the trick,
∂xi
ui = uk δki = uk δik = uk = (uk xi ),k − uk,k xi
∂xk
Hence
Z Z
1 1
< ui > = ui dV = (uk xi ),k − uk,k xi dV
V V V
I Z
1
0
= uk xi nk dS − uk,k xi dV (3.10)
V ∂V V
It is clear that < ui > can not be expressed in terms of boundary data, unless
uk,k = 0.
However, for incompressible materials, such as rubber or plastic zone of
ductile materials, it is often true that uk,k = 0. Therefore,
Z I
1 1
< ui >= ui dV = u0 xi nk dS (3.11)
V V V ∂V k
An average theorem for infinitesimal strain can be stated as follows.
Homogenization I — Classical Averaging Method 25
where 0ij is a constant strain tensor. Then, the average strain field of the RVE
equals the constant strain tensor, i.e.
Proof:
First of all, the prescribed essential boundary condition does not necessarily
generate a constant strain field inside the RVE, i.e.
In fact
ij (x) = 0ij + ˜ij (x), ∀x ∈ V
One may also show the following identities about average virtual work and
average strain energy density.
I
1
< σ : δ >= t · δudS (3.15)
V ∂V
Z Z
1 1
σij δij dV = σij δui,j dV
V V V V
Z I
1 1
= σij δui dV = σij δui nj dS
V V ,j V ∂V
I
1
= ti δui dS (3.17)
V ∂V
σ = σ0 + σd, = 0 + d .
28 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
The stress and strain distributions inside the inhomogeneous solid are
C : (0 + d )
x∈M
σ =
CΩ : (0 + d ) x ∈ Ω
D : (σ 0 + σ d )
x∈M
= Ω 0 d (3.19)
D : (σ + σ ) x ∈ Ω
The Eshelby’s equivalent eigenstrain homogenization method is to choose a
suitable strain field,
0, ∀x ∈ M
= (3.20)
∗ , ∀x ∈ Ω
to superpose with the actual strain field, = 0 + d , such that the total strain
field of homogenized solid is equivalent to the total strain field of inhomoge-
neous solid, i.e.
σ(x) = C : ((x) − ∗ (x))
C : (0 + d ) C : (0 + d ),
x∈M
= 0 d ∗ = (3.21)
C : ( + − ) CΩ : (0 + d ), x ∈ Ω
Consider 0 = D : σ 0 . Under the chosen traction boundary condition, <
σ >= σ 0 , but 0 6=< >.
From (3.21), one may derive that
σ d (x) = C : (d (x) − ∗ (x)), ∀x ∈ V (3.22)
CΩ (0 + d ) = C : (0 + d − ∗ ), ∀x ∈ Ω (3.23)
where Eq.(3.23) is called "stress consistency condition". It is the criterion for
choosing suitable eigenstrain field. Note that 0 + d − ∗ is the total elastic
strain.
Alternatively, Eqs (3.21) and (3.22) can be recast into following forms,
σ = C : ( − ∗ ) ⇒ = D : σ + ∗ (3.24)
σ d = C : (d − ∗ ) ⇒ d = D : σ d + ∗ (3.25)
Figure 3.2. Illustration of Eshelby’s equivalent eigenstress principle. (a) Initial heterogeneous
body, (b)equivalent homogeneous body (V = Ω ∪ M ).
∗ + D : σ ∗ = 0, or σ ∗ + C : ∗ = 0 (3.33)
Denote the average strain and stress in the matrix and in the inhomogeneity
as
Z Z
1 1
¯M := (x)dV , σ̄ M := σ(x)dV ; (3.37)
M M M M
Z Z
1 1
¯Ω := Ω
(x)dV , σ̄ := σ(x)dV ; (3.38)
Ω Ω Ω Ω
D̄ : σ 0 − f DΩ : σ̄ Ω = D : σ 0 − f D : σ̄ Ω (3.42)
Therefore,
D − D̄ : σ 0 = f D − DΩ : σ̄ Ω = f D − DΩ :< σ 0 + σ d >Ω (3.43)
The equaqtion is often referred to as The Basic Equation for Average Stress.
By definition,
σ̄ Ω = CΩ :< 0 + d >Ω (3.44)
32 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
where AΩ = (C − CΩ )−1 : C.
If one can relate the perturbed strain with the eigenstrain, i.e.
d = SΩ : ∗ (3.46)
Subsequently,
Therefore,
Consider
DΩ − D : CΩ = 1(4s) − D : CΩ
and
−1 −1
AΩ = (C − CΩ )−1 : C = C−1 : (C − CΩ )
= 1(4s) − C−1 : CΩ
= 1(4s) − D : CΩ = DΩ − D : CΩ
−1
⇒ DΩ − D : CΩ = AΩ (3.52)
Homogenization I — Classical Averaging Method 33
Therefore −1
D̄ − D : σ 0 = f AΩ − SΩ : D : σ0 (3.53)
It is the straightforward to derive
D̄ = 1(4s) + f (AΩ − SΩ )−1 : D
(3.54)
Note that the crucial step of this derivation is the assumption that disturbance
strain field can be related to eigenstrain distribution, i.e. d = SΩ : ∗ , where
the tensor SΩ is called the Eshelby tensor. Chapter 6 will be devoted to derive
Eshelby tensor for specific shapes of inhomogeneities or inclusions.
His work was a great part of his life. His general field was the theoretical
physics of the deformation, strength and fracture of engineering materials, and
his principal interests were lattice defects and continuum mechanics.
Though motivated by the desire to understand he kept a firm eye on appli-
cation and had no time for useless erudition, like willard Gibbs his object was
to make things appear simple by "looking at them in the right way". With
a keen discrimination he selected those worthwhile difficult problems whcich
nevertheless had some chance of solution. Entirely unconcerned with personal
advancement, he hoped only of his paper that each would be a "little gem".
And so it is. Many indeed are treasure houses, abounding in undeveloped
asides on which others may later build, for often he did not elaborate. He
regarded himself as a modest "supplier of tools for the trade", and he felt to
others their day to day use. His colleagues everywhere were always consulting
him.
Eshelby was elected a Fellow of the Toyal Society in 1974, being "distin-
guished for his theoretical studies of the micromechanics of crystalline imper-
fections and material inhomogeneities". he made major contributions to the
theory of static and moving dispocations and of point defects. By an elegant
use of the theory of the potential he obtained some remarkable results on the
elastic fields of ellipsoidal inclusions and inhomogeneities.
In 1951 he introduced, in analogy with the Maxwell tensor, the elastic en-
ergy momentum tensor, which yields forces on elastic singularities. During
his later years he was much concerned with this concept and its developments,
which can provide parameters characterizing the singular fields.
In 1968 he published accounts of its application to the calculation of forces
on static and moving cracks inelastic media. Related work, formulated for
application also to plastic-elastic media, was published simultaneously and in-
Homogenization I — Classical Averaging Method 35
3.9 Exercises
Probelm 3.1 Let
1 x·x
w(x) = exp(− 2 ), (3.55)
πR3 R
representing a Gaussian distribution .
For any smooth vector field, A ∈ IR3 , define weighted average operation,
Z
< A > (x) := 3
w(x − x0 )A(x0 )dΩx0 (3.56)
IR
(Hint: Use Gauss theorem (divergence theorem), and the fact that w(x) →
0 as |x| → ∞.)
show:
SΩ + D : TΩ : C = 1(4) (3.62)
TΩ + C : SΩ : D = 1(4) (3.63)
where SΩ and TΩ are the Eshelby tensor and the conjugate Eshelby tensor
respectively.
Hint: First show that
σ d = TΩ : σ ∗ , ∀x ∈ Ω (3.65)
where TΩ is the so-called conjugate Eshelby tensor. Show that the effective
elastic tensor is equal to
h i
C̄ = 1(4s) + f (BΩ − TΩ )−1 : C (3.66)
Probelm 3.5 Suppose that an RVE (V) is subjected the following pure trac-
tion boundary condition,
n · σ = t̄ = n · σ 0 , ∀x ∈ ∂V (3.67)
Show that
< σ : δ >= σ 0 :< δ > . (3.68)
Green’s function and Fourier transform 37
Chapter 4
where K is the so-called kernel function. Once the kernel function is deter-
mined, the inverse operator L−1 is determined.
Suppose that we have already known the inverse operator of L in Eqs.(4.3)
and (4.4). We then can solve the differential equation by applying the inverse
operation,
one may find that LK(x − y) = δ(x − y). Therefore, one can deduce that the
kernel function of a differential operator L is its Green’s function:
In principle, if the Green’s function of a BVP has been found, the BVP is
considered to be solved. This is becaruse one can obtain the general solution
Green’s function and Fourier transform 39
Consider the fact that both u(x) and G(x, y) satisfy the same homogeneous
essential boundary conditions. A simple reciprocal holds
Z l Z l
uL(G)dx = GL(u)dx (4.18)
0 0
which leads to
Z l Z l
u(y)δ(x − y)dy = G(x − y)f (y)dy (4.19)
0 0
and consequently,
Z l
u(x) = G(x − y)f (y)dy (4.20)
0
40 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Subsraction of (4.22) from (4.23) yields the so-called Green’s reciprocal theo-
rem:
Z Z n
∂v ∂u o
u∇2 v − v∇2 u dΩ = u − v dS (4.24)
Ω ∂Ω ∂n ∂n
Let v(x) = G(x, y), f1 (x) = f (x), and f2 (x) = δ(x − y). We can then show
that
Z n Z
∂G ∂u o
u(x) = u−G dSy + G(x, y)f (y)dΩy (4.25)
∂Ω ∂n ∂n Ω
Note that in 4.25, the Green’s function solution does not necessarily have the
same boundary data as unknown function, u(x), as in the previous example.
Often times, the Green’s function in the infinite domain is chosen in a recipro-
cal representation.
Lemma 4.3 (Jordan) Suppose that on the circular arc CR shown in Fig.(4.2)
we have f (ξ) → 0 uniformly as R → ∞. Then
We note that if x < 0 similar result holds for the contour in lower half space.
Now, the question becomes what is a residue and how to calculate it. The
answer involves with the singularity of f (z). For a function of complex varible,
f (z), one may express f (z) in a local region by its Laurent expansion – an
extension of Taylor expansion of real variable. For instance around a fixed
point zj , we may write
∞
X ∞
X
f (z) = an (z − zj )n + a−n (z − zj )−n , 0 < |z − zj | < a (4.31)
n=0 n=1
The simple pole at x is only counted for half of the residue is because that it
has only half circle.
Proof:
by definition,
Z ∞ Z ∞h 1 Z ∞ i
f¯(ξ)ḡ(ξ) exp(iξx)dξ = ḡ(ξ) f (y) exp(−iξy)dy exp(iξx)dξ
−∞ 2π −∞
Z−∞∞ h 1 Z ∞ i
= f (y) ḡ(ξ) exp(iξ(x − y))dξ dy
−∞ 2π −∞
Z ∞
1
= g(x − y)f (y)dy (4.38)
2π −∞
In 3D, we have
Z ∞ Z ∞
1
f¯(ξ)ḡ(ξ) exp(iξ · x) = g(x − y)f (y)dy (4.39)
−∞ (2π 3 ) −∞
44 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Hence Z ∞
1
1̄(ξ) = δ(ξ) = exp(−iξx)dx (4.46)
2π −∞
Green’s function and Fourier transform 45
1 d d 0
Use the polar coordinate ∇2 = r and denote x = x − y. We have
r dr dr
1 d d 0 0
r G = −δ(x1 )δ(x2 ) (4.53)
r dr dr
and
Z 2π Z r Z
1 d 0 d 0 0 0 0 0 0
r 0 G r dr dθ = − δ(dx1 )δ(dx2 )dx1 dx2 (4.54)
0 0 r0 dr0 dr Ω
0 p
where r is the dummy variable and r = |x−y| = (x1 − y1 )2 + (x2 − y2 )2 .
The integration domain is a circular region centered at x = y and with the
radius r.
Therefore,
d d 1 1
2π r G = −1, ⇒ G = − (4.55)
dr dr 2π r
Finally, we find that
1
G(x − y) = − lnr (4.56)
2π
Example 4.14 Consider one dimensional Helmhotz equation,
d2 u
+ k 2 u = δ(|x − y|) (4.57)
dx2
Apply Fourier transform,
Z ∞
1
ū(ξ) = u(x) exp(−iξx)dx
2π −∞
d2 u Z ∞ 2
1 d u
F = exp(−iξx)dx = −ξ 2 ū(ξ)
dx2 2π −∞ dx2
Z ∞
1 1
δ̄(|x − y|) = δ(x − y) exp(iξx)dξ = exp(−iξy) (4.58)
2π −∞ 2π
and
1 1
ū(ξ) = exp(−iξy) (4.59)
2π k 2 − ξ 2
Green’s function and Fourier transform 47
Therefore,
Z ∞
u(x) = ū(ξ) exp(iξx)dξ
−∞
Z ∞
1 1
= exp(iξ(x − y))dξ
2π −∞ (k + ξ)(k − ξ)
2
exp(iξ(x − y))
I
1 iX
= dξ = Residues of ξ at ξi
2π R (k + ξ) k − ξ) 2
i=1
i n 1 1 o
= − exp(ik(x − y)) − exp(−ik(x − y))
2 2k 2k
i n o
= − cos k(x − y) + i sin k(x − y) − cos k(x − y) − i sin k(x − y)
4k
i 1
= − 2i sin k(x − y) = sin k(x − y) (4.60)
4k k
1 1
δ̄(x − x0 ) = exp −iξ · x 0
⇒ δ(x − x0
) = exp iξ · (x − x0
) dξ
π3 2π 3
Z ∞
0
G(x − x ) = Ḡ(ξ) exp iξ · x dξ
−∞
Z ∞
0
G,ii (x − x ) = − Ḡ(ξ)ξi ξi exp(iξ · (x − x0 ))dξ (4.62)
−∞
and
1 1 1
Ḡ(ξ)ξi ξi = ⇒ Ḡ(ξ) = (4.63)
(2π)3 (2π)3 ξi ξi
Hence
Z ∞
0 1 sin ξr
G(x − x ) = 2
dξ
2π 0 ξr
Z ∞
1 sin ξr 1
= 2
d(ξr) = 2 Si(∞) (4.67)
2π r 0 ξr 2π r
Z x
sin t π
where Si(x) := dt and Si(∞) = . Finally, we have
0 t 2
0 1 1
G(x − x ) = (4.68)
4π |x − x0 |
We let
G∞ ∞
σij m = Cijkl G
kl (4.71)
fim = δ(x − y)δmi (4.72)
where δ(x − y) := δ(x1 − y1 )δ(x2 − y2 )δ(x3 − y3 ), and the integer m is a
free index, which indicates the direction of the concentrated load.
Then,
G∞ ∞ ∞ G∞∞
σij m = Cijkl G
kl = Cijkl Gmk,l → σij,j = Cijkl Gmk,lj
m
Then Green’s function for an infinite linear elastic medium is the solution of
the following equatin,
Cijkl G∞
mk,lj + δ(x − y)δmi = 0 (4.73)
Figure 4.4. The unit sphere S 2 in the ξ-space. Green’s function at point z is expressed by a
line integral along S 1 which lies on the plane perpendicular to z
Let z = x − y. We have
Z ∞
∞ 1 1
2
Gij (z) = (λ+2µ)δ ij ξ −(λ+µ)ξ i ξj exp(iξ·z)dξ
(2π)3 −∞ µ(λ + 2µ)ξ 4
(4.86)
2
To integrate (4.86), we donote S as a unit sphere where |ξ| = 1, and denote
S 1 as a unit circle on the surface of S 2 , where S 2 is intersected by a plane
perpendicular to vector z.
Apply Radon decompositon,
where ξ 2 = ξ12 + ξ22 + ξ32 and dS is the surface element on the unit sphere S 2
in ξ-space. Imagine that the ξ-space is a expanded spherical balloon.
Denote ξ̄ = ξ¯i eξi as a unit vector pointing from the origin to the surface of
2
S along ξ direction and denote z̄ = z̄i ezi as another unit vector point from the
origint to thepsurface of S 2 along z direction. Therefore, ξ = ξ ξ̄ and z = zz̄
where ξ = ξ12 + ξ22 + ξ32 and z = z12 + z22 + z32 . Obviously, ξ¯i = ξi /ξ
p
since Z ∞
exp(iξz ξ̄ · z̄)dξ = 2πδ(z ξ̄ · z̄) (4.92)
−∞
one has
[(λ + 2µ)δij − (λ + µ)ξ¯i ξ¯j ]
I
1
Gij (z) = δ(z ξ̄ · z̄) dS(ξ̄) (4.93)
2(2π)2 S2 µ(λ + 2µ)
To integrate (4.93), one has to evaluate the following two integrals:
Z Z
δ(z ξ̄ · z̄)dS? and ξ¯i ξ¯j δ(z ξ̄ · z̄)dS?
S2 S2
Green’s function and Fourier transform 53
Consider ξ̄ · z̄ = cos θ, d cos θ = − sin θdθ. One may decompose the sur-
face element
on S 2 into:
dS(ξ̄) = sin θdθdφ = −d(ξ̄ · z̄)dφ, where θ →
[0, π] cos θ → [1, −1] and φ → [0, 2π]. If we let t = ξ¯ · z̄,
Z Z 1 Z 2π
2π
δ(z ξ̄ · z̄)dS = δ(zt)dt dφ = (4.94)
S2 −1 0 z
On the other hand,
Z Z 1 Z 2π
ξ¯i ξ¯j δ(z ξ̄ · z̄)dS = δ(zt)ξ¯i ξ¯j dtdφ (4.95)
S2 −1 0
Considering,
a1 = (a1 · ei )ei ; a2 = (a2 · ei )ei
one has
Thereby,
where t = cos θ.
54 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Example 4.16 Consider an image density function, g(x, y). The two-dimensional
Radon transform may be defined as
Z ∞Z ∞
ĝ(ρ, θ) = g(x, y)δ(ρ − x cos θ − y sin θ)dxdy (4.106)
−∞ −∞
(a)
(b) (c)
Figure 4.6. Two-dimensional Radon transform: (a) prjectinline, (b) image in the physical
space, and (c) image in the Radon transform space
Example 4.17 Let f (x) = δ(x). The Radon transform of Dirac’s delta
function is
Z ∞
δ̂(s, n) = R(δ) = δ(x)δ(s − n · x)dx = δ(s) (4.109)
−∞
where s = ni xi .
Subsequently,
δ̃(s, n) = δ 00 (s) (4.110)
and the inverse Radon transform is
Z
1
δ(x) = − δ 00 (nk xk )dS (4.111)
8π 2 S2
Figure 4.7. Two-dimensional Radon transform: (a) projection line, (b) image in the physical
space, and (c) image in the Radon transform space
∂2
Applying the harmonic operator ∇2 = to the above identity and con-
∂xi ∂xi
sidering Example (4.15) (Eq.(4.68)) yields
Z 1
δ 00 (nk xk )ni ni dS = 2π∇2 = −8π 2 δ(x) (4.113)
S2 |x|
Now we use the Radon transform to derive 3D static Green’s function of a
linear elastic medium. Consider the concentrated load is acting at the origin of
the coordinat (y = 0).
Assume that the Green’s function can be written as a form of inverse Radon
transform, Z
∞ 1
Gkm (x) = − 2 G˜∞ (ξ¯n xn )dS (4.115)
8π S 2 km
Then Z
∞ 1 00
Gkm,lj (x) = − 2 G˜∞
km (ξ¯n xn )ξ¯l ξ¯j dS (4.116)
8π S 2
On the other hand,
Z
−1
1
δ(x) = R δ̃(s) = − 2 δ 00 (ξ¯n xn )dS (4.117)
8π S 2
We then obtain
00 00 ¯
Cijkl ξ¯j ξ¯l G˜∞ ¯
km (ξn xn ) = −δim δ (ξn xn ) (4.118)
58 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
which leads to
¯ −1 ¯ ¯
G˜∞
ij (ξ) = −Kij (ξ)δ(ξn xn ) + C1 ξn xn + C0 (4.119)
where
¯
Nij (ξ)
−1
Kri Cijkl ξ¯l ξ¯j = δrk , or Kij = (4.120)
D(ξ)
Note that C1 = C0 = 0 because it is required that G∞
ij (x) → 0, as x → ∞.
For isotropic materials,
−1 1h (λ + µ)ξ¯i ξ¯k i
Kij (ξ̄) = δij − (4.121)
µ (λ + 2µ)
and, correspondingly,
Z
1 −1
G∞
ij (x) = 2 Kij (ξ̄)δ(ξ¯n xn )dS (4.122)
8π S2
and subsequently,
1 h δij (λ + µ) i
G∞
ij (x) = − |x|,ij (4.123)
4πµ |x| 2(λ + 2µ)
would remain a marked man. No one had been executed in Auxerre but Fourier
had been an agent of the terror there. His arrest was on a charge of H’ebertism
and the H’ebertists were to the left of Robespierre. The word ’terrorist’ then,
like ’Trotskyist’ now, denoted a defeated yet feared opponent.
Luckily an opportunity to leave Auxerre now presented itself. A new col-
lege (the Echole Normale) was being set up in Paris to help train teachers and
Fourier could now study under men like Lagrange, Monge and Laplace and
excape his terrorist past. Fourier’s talents were soon noted, but the college was
not successful and its closure was followed by further problems for Fourier.
’We shudder when we think that the pupils of the Ecole Normale were cho-
sen under the reign of Robespierre and his proteges. It is only too true that
Balme and Fourier, pupils of the department of Yonne have long prefessed the
atrocious principles and infernal maxims of the tyrants. Nevertheless they pre-
pare to become teachers of our children. Is it not to vomit their poison in the
bosim of innocence (From an address to the National Convention, quoate by
Herivel)’
60 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Fourier was again arrested, released, rearrested and finally, following yet
another political swing, released to become a teacher at the new Ecole Poly-
technique.
Here Fourier remained for three years. That his talent was recognized is
shown by the fact that he succeeded Lagrange in the Chair of Analysis and
Mechanics. The quiet interlude was ended by a gonernment order to join the
invasion of Egypt. Ostensible intended to liberate Egypt from the Turks and to
threaten the British position in India, the expedition may have been seen by the
government as a way of keeping a troublesome general as far away as possible
and by the general (Napoleon) as the first step toward becoming Emperor of
the East. Fourier wa one of a ghroup of scientists and intellectualls intended to
form part of the immense cultural benefits that France was to bestow on Egypt.
Both before and after Napolean’s departure, Fourier occupied several im-
portant administrative and political posts in Egypt. When the French expedi-
tion finally surrended in 1801 and Fourier was repatriated, Napoleon offered
him the post of Prefect of the Department of the Isere centred round Grenoble
(France had been divided into 83 Departments and each Prefect governed his
Department of behalf of the central government.)
Although he could have continued a Professor at the Polytechnique, Fourier
accepted the offer. Herivel suggests that Egypt had given him a taste for admin-
istration and that he hoped to rise higher. Herivel also accounts that Fourier’s
close association with Kleber after Napoleon’s departure account for the fact
that these hopes were not fulfilled.
Fourier seems to have been popular and efficient Prefect. His greatest achieve-
ment during his 14 years of office was by reconciling the conflicting interests
of some forty communities to enable the swamps of Bourgion to be drained.
The draining of twenty thousand acres of swamps resulted in major economic
and health benefits and was achieved during a period morenoted for grandiose
paper plans than for concrete achievements. Fourier’s other administrative
memorial was a new road across the Alps (now Route 91).
Apart from his perfectorial duties Fourier helped organize the Description
of Egypt. This work written by the intellectuals attached to the Egyptian ex-
pedetion did much to inspire European interest in Egypt and was thus one of
the two permanent results of the expediton. (The other was the discovery of
the Rosetta Stone, atrilingual inscription which was to provide the key to the
deciphering of hieriglyphics.)
Fourier’s main contribution was the general introduction – a survey of Egyp-
tian history up to modern times. An Egyptologist with whom I discussed this
described the introduction as a masterpiece and a turning point in the subject,
was surprised to hear that Fourier also had a reputation as a mathematician!
4.7 Exercises
Probelm 4.1 Find the Green’s function for a both end clamped Euler-Bernoulli
beam, i.e.
d2 d2 G(x, y)
EI = δ(x − y), ∀x, y ∈ (0, `) (4.124)
dx2 dx2
and
G(0, y) = G(`, y) = 0, G0 (0, y) = G0 (`, y) = 0 . (4.125)
Probelm 4.2 For isotropic materials, elasticity tensor has the form
Show
1.
Kik (ξ) = Cijk` ξj ξ` = (λ + µ)ξi ξk + µδik ξj ξj (4.127)
2. (Hint : use eijk eimn = δjm δkn − δjn δkm .)
1
Nij (ξ) = eik` ejmn Kkm K`n
2
= µξ 2 ((λ + 2µ)δij ξ 2 − (λ + µ)ξi ξj ) (4.128)
3.
D(ξ) = µ2 (λ + 2µ)ξ 6 (4.129)
and
Z ∞ Z ∞
exp(i(ξ1 x1 + ξ2 x2 ))
dξ1 ξ2
−∞ −∞ ξ12 + ξ22
= −2πlnR (4.133)
62 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
q
0 0
where R = (x1 − x1 )2 + (x2 − x2 )2 .
1 δij 1 ∂2
G∞ 0
ij (x, x ) = − |x − x0 | (4.134)
4πµ |x − x0 | 16πµ(1 − ν) ∂xi ∂xj
Chapter 5
EIGENSTRAIN THEORY
Let,
Z ∞
uk (x) = ūk (ξ) exp(iξ · x)dx
Z−∞
∞
= ūk (ξ) exp(iξm xm )dx (5.5)
Z−∞
∞
∗ k` (x) = ¯∗k` (ξ) exp(iξm xm )dx (5.6)
−∞
Hence
Z ∞
uk,`j (x) = − ūk ξ` ξj (ξ) exp(iξm xm )dx (5.7)
−∞
Z ∞
∗
k`,j (x) = i ¯∗k` (ξ)ξj exp(iξm xm )dx (5.8)
−∞
which leads to
Cijk` ξj ξ` ūk = −iCijk` ¯∗k` (ξ)ξj (5.10)
Denote
f¯1
K11 K12 K13 ū1
K21 K22 K23 ū2 = f¯2 (5.13)
K31 K32 K33 ū3 f¯3
We find that
Nij (ξ) ¯ −1 ¯
ūi (ξ) = fj = Kij fj (5.14)
D(ξ)
where
1
Nij (ξ) = eik` ejmn Kkm K`n (5.15)
2
D(ξ) = emn` Km1 Kn2 K`3 (5.16)
Eigenstrain Theory 65
subsequently,
h i
A(λ + 2µ)|ξ|2 + B(λ + µ) ξ ⊗ ξ + Bµ|ξ|2 1(2) = 1(2) (5.21)
−1 |ξ|−2 n (λ + µ) o
Qij = Kij = − ξi ξj + δ ij (5.25)
µ µ(λ + 2µ)|ξ|2
Consider
Nij (ξ)
ūi (ξ) = Qij (ξ)f¯j = −iCj`mn ¯∗mn ξ` (5.26)
D(ξ)
Applying Fourier inverse transform,
Z ∞
Nij (ξ)
ui (x) = −i Cj`mn ¯∗mn (ξ)ξ` exp(iξ · x)dξ (5.27)
−∞ D(ξ)
Z ∞
Nij (ξ)
= − f¯j (ξ) exp(iξ · x)dξ (5.28)
−∞ D(ξ)
66 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
(a) (b)
Figure 5.1. Illustraions of dislocations: (a)edge dislocation, and (b) screw dislocation
(λ + µ) ξ2 ξ3
Q23 (ξ) = −
µ(λ + 2µ) ξ 4
Q32 (ξ) = Q23 (ξ)
[(λ + 2µ)ξ 2 − (λ + µ)ξ32 ]
Q22 (ξ) = (5.54)
µ(λ + 2µ)ξ 4
Obviously,
Z ∞
δ(ξ3 )Q12 (ξ)ξ3 dξ3 = 0
Z−∞
∞
δ(ξ3 )Q13 (ξ)ξ2 dξ3 = 0
Z−∞
∞
δ(ξ3 )Q22 (ξ)ξ3 dξ3 = 0
Z−∞
∞
δ(ξ3 )Q23 (ξ)ξ2 dξ3 = 0
Z−∞
∞
δ(ξ3 )Q32 (ξ)ξ3 dξ3 = 0
Z−∞
∞
1 ξ2
δ(ξ3 )Q33 (ξ)ξ2 dξ3 =
−∞ µ (ξ12 + ξ22 )
Thereby, u1 (x) = u2 (x) = 0, and
Z ∞Z ∞
b ξ2
u3 (x) = exp i(ξ1 x1 + ξ2 x2 ) dξ1 dξ2
(2π)2 −∞ −∞ ξ1 (ξ12 + ξ22 )
b x
2
= tan−1 (5.55)
π x1
according to the inverse Fourier transform (Mura’s book page 17),
Z ∞Z ∞
ξ2
−1 x2
2 2 exp i(ξ x
1 1 + ξ x
2 2 ) dξ dξ
1 2 = 2π tan
−∞ −∞ ξ1 (ξ1 + ξ2 ) x1
For simplicity, we assume that the crack opening has the following form:
s
x21 x22
[u3 ] = b 1− − 2 χ(Ω) (5.57)
a21 a2
where parameter b is the Burger’s vecter, and χ(Ω) is the characteristic func-
tion of crack region, which can be defined as interpreted as
1, ∀x ∈ Ω
χ(Ω) = H(Ω − x) = (5.58)
0, ∀x ∈ IR3 /Ω
Therefore,
Z Z Z ∞ 0
∗ (x0 ) exp −iξ · (x − x0 dx =
−∞
s
x02 x02
Z Z
0
b 1− 1
− 2
exp(−iξ 3 x3 − iξ · (x − x ) dx01 dx02 (5.60)
Ω a21 a22
where in the second line, all vectors become 2D vectors, i.e. ξ = ξ1 e1 + ξ2 e2
and x = x1 e1 + x2 e2 .
Employ the fundamental formula of micro-elasticity,
Z ∞ Z ∞
i
ui (x) = Cj`mn ∗ (x0 )ξ` Nij (ξ)D−1 (ξ)
(2π)3 −∞ −∞
o
exp −iξ · (x − x0 ) dξ dx0 (5.61)
and
s
Z ∞Z 0
b x0 2 x22 ξq ξ` Nkp (ξ)
uk,` (x1 , x2 , 0) = Cpq33 1 − 2 − 2
(2π)2 −∞ Ω a1 a2 D(ξ)
0
· exp −iξ · (x − x )dΩx0 dξ (5.63)
subsequently,
Z ∞
b Cijk` Cpq33 ξq ξ` Nkp (ξ)
σij = Cijk` uk,` =
(2π)3−∞ D(ξ)
Z s 02 02
x x 0
0 0
1 − 2 + 2 exp −iξ · (x − x ) dx1 dx2 (5.64)
Ω a1 a2
We first calculate the inverse Fourier transform along ξ3 , i.e. evaluating the
following integral,
Z ∞
Cijk` Cpq33 ξq ξ` Nkp (ξ) 0
exp(−ξ · (x − x )dξ3 . (5.65)
−∞ D(ξ)
For isotropic materials,
Nkp (ξ) [(λ + 2µ)δkp ξ 2 − (λ + µ)ξk ξp ]
= (5.66)
D(ξ) µ(λ + 2µ)ξ 4
74 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Since the problem is symmetric, we only consider the upper halp space (x3 >
0). Because the convergence requirement of Fourier transform, we are only
interested in the root with a negative imaginary part, i.e.
q
ξ3N = −i ξ12 + ξ22 (5.68)
p ( 0
)
2π
y 0 1 − y 02 dy
∂2
Z r Z
ibµ(λ + µ) η1 2 η2 2 0 0
σ33 = + 0 dy1 dy2
x3 =0 2π(λ + µ) 0 a1 Ω g + y cos(θ − φ)
a2 ∂g 2
( Z 1 0p 0
)
ibµ(λ + µ) 2π η1 2 η2 2 ∂ 2
r
y 1 − y 02 dy
Z
= + (5.80)
∂g 2 0
p
2π(λ + µ) 0 a1 a2 g2 − y02
Let p 0
1
y0 1 − y 02 dy
Z
I= p
0 g2 − y02
Change of variable
0 g2 − 1 1 2
y2 =1− w− (5.81)
4 w
One can show that
∂2I 1 g+1 g
= − ln + . (5.82)
∂g 2 2 g − 1 g2 − 1
76 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
ibµ(λ + µ) 2π η1 2 η2 2 gdφ
Z r
σ33 (x1 , x2 , 0) = + + O(1) (5.88)
2π(λ + µ) 0 a1 a2 g 2 − 1
Let
η2 η22 1/2 y
1
f (η) = g + , and ŷ = . (5.89)
a21 a22 y
Eigenstrain Theory 77
Figure 5.4. The shortest distance between the crack surface and a point
ibµ(λ + µ) 2π f (η − f (ŷ)
Z Z 2π
1 f (ŷ)
σ33 (x1 , x2 , 0) = dφ + dφ
2π(λ + µ) 0 g2 − 1 2π 0 g 2 − 1
Z 2π
1 f (ŷ)
= dφ + O(1) (5.90)
2π 0 g 2 − 1
Assume that g = −η · y = y cos ψ. Then
Z 2π Z 2π
dφ d(φ − ψ) −2π 2iπ
2 2 2
=p =p . (5.91)
0 g − 1 0 y cos (φ − ψ) − 1 1−y 2 y2 − 1
and
bµ(λ + µ) ŷ · y ŷ1 ŷ2 1/2
σ33 (x1 , x2 , 0) = + (5.92)
(λ + 2µ) y 2 − 1 a21 a22
p
y→ŷ
where
`22
`2 `23
1
g := ++ (5.105)
a21
a22 a23
x ` x2 `2 x3 `3
1 1
f := + + 2 (5.106)
a21 a22 a3
x2 x2 x2
e := 1 − 21 + 22 + 23 (5.107)
a1 a2 a3
Eq. (5.142) has two roots,
f f2 e 1/2
r(`) = − ± 2 + (5.108)
g g g
f2 e 1/2
Since + is even in `, while gimn (`) is odd in `,
g2 g
Z 2
f e 1/2
2
+ gimn (`)dw = 0 (5.109)
S2 g g
Let λ1 = `1 /a21 ,λ2 = `2 /a22 and λ3 = `3 /a23 . We have
∗ mn
I
f
ui (x) = gimn (`)dw
8π(1 − ν) S 2 g
∗ mn
I
x` λl
= gimn (`)dw
8π(1 − ν) S 2 g
∗ mn x`
I
λ`
= gimn (`)dw (5.110)
8π(1 − ν) S 2 g
Then
∗ mn δ`j
I
λ`
ui,j (x) = gimn (`)dw
8π(1 − ν) S 2 g
∗ mn
I λ
j
= gimn (`)dw (5.111)
8π(1 − ν) S 2 g
One can find induced elastic strain field by symmetrizing the elastic distor-
tion,
∗ mn
I
1 λi gjmn + λj gimn
ij = (ui,j + uj,i ) = dw (5.112)
2 16π(1 − ν) S 2 g
`i
where λi = is the component of the normalized vector λ = λi ei and
a2i
g = λ · λ = λ2 .
Eigenstrain Theory 81
Consider
The last two indices of the third order tensor gijk is symmetric. We can then
define a fourth order symmetric tensor,
I
Ω 1 λi gjmn + λj gimn
Sijmn := dw (5.114)
16π(1 − ν) S2 g
It is obvious that
Ω Ω Ω
Sijmn = Sijnm = Sjimn
where the superscript indicates that the Eshelby tensor is for induced strain
field inside the ellipsoidal, Ω.
Remark 5.5.1 The most amazing fact of this result is that the induced strain
field and stress field inside the inclusion are uniform, and the Eshelby tensor
for any ellipsoidal shape of inclusion is a constant tensor.
82 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
where the upper case indices are not summed with lower case indices.
Ω , we consider
Example 5.2 To compute S1111
Ω
8π(1 − ν)S1111 = 2νI1 (0) + J11 (0) + 2(1 − ν)2I1 (0) + 2J11 (0)
= (4 − 2ν)I1 (0) + 3J11 (0)(a21 I11 (0) − I1 (0))
= (1 − 2ν)I1 (0) + 3a21 I11 (0) (5.120)
which leads to
Ω 3a21 (1 − 2ν)
S1111 = I11 (0) + I1 (0) (5.121)
8π(1 − ν) 8π(1 − ν)
The integral II (0) and IIJ (0) can be expressed in terms of standard elliptic
integrals. For example, assuming a1 > a2 > a3 , we have
4πa1 a2 a3
I1 (0) = {F (θ, k) − E(θ, k)}
(a21 − a22 )(a21 − a22 )1/2
4πa1 a2 a3 n a (a2 − a2 )1/2 o
2 1 3
I3 (0) = − E(θ, k)
(a22 − a23 )(a21 − a23 )1/2 a1 a3
where
Z θ
dt
F (θ, k) =
0 (1 − k 2 sin2 t)1/2
Z θ
E(θ, k) = (1 − k 2 sin2 t)1/2 dt (5.122)
0
Eigenstrain Theory 83
h i1/2
and θ = sin−1 (1 − a23 /a21 )1/2 , k = (a21 − a22 )/(a21 − a23 ) .
In applications, the following invariant formulas are very useful,
When the ellipsoidal becomes a sphere, Eshelby tensor become simple num-
bers. Let a1 = a2 = a3 = a. We have
4π
IIs =
3
s 4π 1
II,J =
5 a2
s 8π
JIJ = −
15
and hence
5ν − 1 2(4 − 5ν)
Ω
Sijk` = δij δk` + (δik δj` + δjk δi` ) (5.123)
15(1 − ν) 15(1 − ν)
where
−1
Cj`mn G∞ 0
ij,` (x − x ) = ·
8π(1 − ν)
n δmi (xn − x0n ) + δni (xm − x0m ) − δmn (xi − x0i )
(1 − 2ν)
|x̄|3
(xm − xm )(xn − xn )(xi − x0i ) o
0 0
+3
|x̄|5
−1 n ∂3 h ∂ δ
mi
= |x̄| − 2(1 − ν)
8π(1 − ν) ∂xi ∂xm ∂xn ∂xn |x̄|
∂ δni i ∂ 1 o
− 2νδmn (5.125)
∂xm |x̄| ∂xi |x̄|
Introduce the following potential functions,
Z
ψ(x) = |x − x0 |dΩx0 (5.126)
ZΩ
1
φ(x) = 0
dΩx0 (5.127)
Ω |x − x |
where ψ(x) is the biharmonic potential, whereas φ(x) is the Newtonian poten-
tial. This is because of the fact
−8π x ∈ Ω
∇4 ψ = 2∇2 φ = (5.128)
0 x ∈ IR3 /Ω
∂2
Z
2
∇ ψ = |x − x0 |dΩx0
∂x2` Ω
(x` − x0` )(x` − x0` ) o
Z n
δ``
= − dΩx0
Ω |x̄| |x̄|3
Z
2
= dΩx0 = 2φ(x) (5.129)
Ω |x̄|
Subsequently,
∇4 ψ = ∇2 ∇2 ψ = 2∇2 φ
Z Z
1 1
= 2 ∇ dΩ = 8π ∇2 dΩ
Ω |x̄| Ω 4π|x̄|
Z
= 8π ∇2 GL (x − x0 )dΩx0 (5.130)
Ω
Eigenstrain Theory 85
∞ 1
Sijk` (x) = ψ,ijk` (x) − 2νδk` φ,ij (x)
8π(1 − ν)
−(1 − ν)(δki φ,`j (x) + δ`i φ,kj (x)
+δkj φ, `i(x) + δ`j φ,ki (x)) (5.137)
where
1
Gijmn (x − y) = − Ck`mn Gik,`j (x − y) + Gjk,`i (x − y)
2
1 h
= (1 − 2ν)(δim δjn + δin δjm − δij δmn )
8π(1 − ν)r3
+3ν(δim `j `n + δin `j `m + δjm `i `n + δjn `j `m )
i
+3δij `m `n + 3(1 − 2ν)δmn `i `j − 15`i `j `m `n (5.148)
where Gijmn is called the fourth order Green’s function (the second derivative
of the Green’s function).
If ∗ mn (x) is constant inside the inclusion, the exterior Eshlby tensor can
be defined as
Z
∞
Ḡijmn (x) := Gijmn (x − y)dΩy = Sijmn (5.149)
Ω
which is the weak discontinuity at the interface between matrix and inculsion.
88 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
he would stand back and survey the board. Agter a few moments, he would
announce something like, "Well, there’s a sign error there. You can correct it
and work through to the result for yourselves." As if.
As time went by, our horror at his teaching style gave way to an understand-
ing that the man was, in fact, a genius. Eccentric, yes but a genius. Apparently
addicted to cheap cigars, he would smoke them down to the smaalest butt, then
draw a cherry pipe out of his pocket, and stuff the remains of the cigar into it,
tob e smoked until not a scrap of tobacco was left. He cared little for what peo-
ple thought of hom, I think, and did not pay much attention to the politics of
academia and the scientific community. This resultd in anunconscionalbe de-
lay in his being elevated to the rank of Fellow of the Royal Society, which does
seem to have been a sore point. In one memorable lecture, he described all of
the current theories on a particular topic, listing the names of their authors on
an uncharacteristically cleared chalkboard. He then described what was wrong
with each of their work, condemning the weak-mindedness of these "so-called
scientists" in quite direct terms. Having disposed of their failed logic, he then
wrote the magical letters "FRS" after each of the names. He was elected an
FRS himself that year and did not repeat the performance as far as I can gather.
Eshelby’s impact on material sience is far, far out of proportion to the num-
bers of his publications. In total, he published less than 20 papers over his
entire career (This is not true by the way. Eshelby published alomost 50-to-
60 papers in his lifetime, but the point is valid: this days, you can see a lot
of mediocre people published hundreds of junks, and good papers can not be
published–Li’s comment), but each of them is a classic. A fine demonstraion
of the futility of today’s obsessiion with publication-counting as a means of
career assessment. Eshelby’s work is characterized by real physical insight,
complemented by elegant mathematical analysis (He was a professor of ap-
plied mathematics at Sheffield, in addition to being a professor of the theory of
materials.) In contrast with his lectures, his written work is a modl of clarity.
Although he was a powerful mathematician, he felt that we should only engage
in "mathematical weightlifting" if we could not reason our way to the desired
result through simple physical logic. Goodness knows what he would have
made of today’s computer simulation techniques. I think he would probably
have thought of them as the last desperate resort after both physical reasoning
and mathematical analysis failed.
An insight into Eshelby’s motivations was provided to us in an informal mo-
ment one day, sitting in the small but splendid museum of glassware belonging
to the Faculty of Materials, in a traditional British tea break. The usually unap-
proachable Esheby was unusually affable that day–perhaps he had just receibed
word of his FRS election–but we fell into conversation and one undergraduate
student adked him what had led to his being a "pure theoretician". He told
us the story of a formative experience in his life. It seems that as a young
90 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
5.8 Exercises
Probelm 5.1 Show that the integral
r
π 3 J3/2 (η)
Z
exp{iξ · x}dVx = 4π a (5.153)
V0 2 η 3/2
p
where V0 is a sphere with radius a; η = a|ξ|, and |ξ| = ξ12 + ξ22 + ξ32 .
Hint:
(1)Consider the identity
∇x exp iξ · x = iξ exp iξ · x
(2) Z 1
t sin(a|ξ|t)dt = Γ(1)(a|ξ|)−1/2 J3/2 (a|ξ|)
0
r
π
where Γ(1) = , J (η) is the Bessel function of the first kind.
2 3/2
Probelm 5.2 Derive the displacement field inside an inclusion in which pre-
scribed eigenstrain is a linear functin of coordinates, i.e. Example 5.1.
Eigenstrain Theory 91
Probelm 5.3 Derive Green’s functin for plane strain problem by solving the
following Navier equations,
σβα,β + δ(x − y)δαγ = 0 (5.154)
where γ is the direction that the concentrated force point at.
Assume that the 2D elastic tensor is
Cαβζη = λδαβ δζη + µ(δαζ δβη + δαη δβζ ), α, β, ζ, η = 1, 2 (5.155)
define 2D permutaion symbol
eαβ : e11 = 0, e12 = 1, e21 = −1, e22 = 0 (5.156)
The corresponding e-δ identities are:
δα1 δα2
(1) eαβ =
δβ1 δβ2
(2) eαζ eβη = δαβ δζη − δαη δβζ
(3) eαη eβη = δαβ (5.157)
(4) eαη eαη = δαα = 2! (5.158)
Hints:
Z ∞ Z ∞
exp(i(ξ1 x1 + ξ2 x2 ))
dξ1 ξ2 = −2π ln R (5.159)
ξ12 + ξ22
Z−∞ −∞
∞ Z ∞
ξα ξβ xα xβ
4
exp(iξ · x)dξ = −πδαβ ln R − π 2 (5.160)
−∞ −∞ ξ R
p
where R = x21 + x22 .
Probelm 5.4 Let Ω be the half plane (x2 = 0, x1 < 0), and ∗21 be pre-
scribed as
b
∗21 (x) = δ(x2 )H(−x1 ) (5.161)
2
Show
b x
2 b 1 x1 x2
u1 (x) = tan−1 + (5.162)
2π x1 4π 1 − ν x21 + x22
where ν is the Poisson’s ratio.
Hint: (Mura’s book page 17)
Z ∞Z ∞
ξ2 −1 x2
exp{i(ξ1 x1 + ξ2 x2 )}dξ1 dξ2 = 2π tan
ξ1 (ξ12 + ξ22 ) x1
Z−∞∞ Z−∞∞
ξ1 ξ2 πx1 x2
2 2 2 exp{i(ξ1 x1 + ξ2 x2 )}dξ1 dξ2 = − x2 + x2
−∞ −∞ (ξ1 + ξ2 ) 1 2
92 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
x21 x22
where ∗αβ is a constant tensor, and Ω := x + 2 ≤1 .
a21 a2
Find the Eshelby tensor for interior problem (x ∈ Ω). Hint (see Li (2000)
pages 5606-5607 ).
4πa2
I
`i `j dS = δij (5.167)
S2 3
4πa2
Z
`i `j `m `n dS = (δij δmn + δim δjn + δin δjm ) (5.168)
S2 15
to show that
I
Ω 1 λi gjmn + λj gimn
Sijmn = dS
16π(1 − ν) S 2 g
5ν − 1 2(4 − 5ν)
= δij δmn + (δim δjn + δjm δin ) (5.169)
15(1 − ν) 15(1 − ν)
where Gijmn is called the fourth order Green’s function (the second derivative
of the Green’s function).
Effective Elastic Modulus 95
Chapter 6
We now present Eshebly’s equivalent eigestrain theory and its related engi-
neering homogenization methods.
Z Z
1 1
< σ >M := σdV, < >M := dV (6.1)
M M M M
Z Z
1 1
< σ >α := σdV, < >α := dV (6.2)
Ωα Ωα Ωα Ωα
96 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
By definition,
Z Z
1 1
<σ> = σdV = σdV
V V V M ∪Ωα
Z n Z
1 hM X Ωα i
= σdV + σdV
V M M Ω
α=1 α Ωα
X
= f0 < σ >M + fα < σ >α (6.3)
α
Therefore,
X
f0 < σ >M = <σ>− fα < σ >α
α
X
= C̄ :< > − fα Cα :< >α (6.4)
α
Therefore,
X
f0 < >M = <>− fα < >α
α
X
= D̄ :< σ > − fα Dα :< σ >α (6.9)
α
and
hM 1 Z i
f0 < >M = f0 D :< σ >M = D : σdV
V M M
h1 Z X Ωα 1 Z i
= D: σdV − σdV
V V α
V Ωα Ωα
X
= D: <σ>− fα < σ >α (6.10)
α
We name Eqs. (6.6) and (6.11) as the basic equations of average stress/strain
fields.
t = n · σ 0 , ∀ x ∈ ∂V
u(x) = x · 0 , x ∈ ∂V
0 =< > .
98 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Since by definition < σ >α = Cα :< >α and < >α = Dα :< σ >α , one
can rewrite Eqs. (6.27) and (6.28) as
α
C : Aα : D : σ 0
< σ >α = (6.31)
Bα : σ0
or
Dα : B α : C : 0
< >α = (6.32)
Aα : 0
Suppose that prescribed macro-stress boundary condition is applied. Sub-
stituting both expressions in Eq. (6.31) into the basic average equation (6.11)
yields,
α
Xn C : Aα : D̄ : σ 0
(D̄ − D) : σ 0 = fα (Dα − D) : (6.33)
α
α=1 B : σ0
Finally, we obtain
n
X
fα (Aα − Sα )−1 : D
D +
α=1
D̄ = (6.36)
n
X
fα D : (Bα − Tα )−1
D −
α=1
100 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Note that the index α starts from 1, and each α is an inhomogeneous phase.
One of the drawback of dilute distribution homogenization is
D̄ : C̄ 6= 1 or D̄ 6= C̄−1 .
Obviously, the effective elastic stiffness is not consistent with the effective
compliance tensors.
Effective Elastic Modulus 101
Aα = (C − Cα )−1 : C
Here C is the elastic tensor of the matrix, which is assumed to be isotropic, i.e.
C = 3KE(1) + 2µE(2) . We can then calculate
and
Aα = (C − Cα )−1 : C
1 (1) 1 (2)
(1) (2)
= E + E : 3KE + 2µE
3(K − K α ) 2(µ − µα )
K µ
= α
E(1) + E(2)
K −K µ − µα
Since the composite is isotropic, we use the Eshelby tensor of spherical inclu-
sions, For spherical inclusion, the Eshelby tensor is
5ν − 1 (2) 2(4 − 5ν) (4s)
SΩ = 1 ⊗ 1(2) + 1
15(1 − ν) 15(1 − ν)
(1 + ν) (1) 2(4 − 5ν) (2)
= E + E
3(1 − ν) 15(1 − ν)
= s1 E(1) + s2 E(2)
1+ν 2(4 − 5ν)
where s1 = and s2 = .
3(1 − ν) 15(1 − ν)
Then
K µ
Aα − Sα = − sα
1 E(1)
+ − sα
2 E(2)
(K − K α ) (µ − µα )
and
−1 1 1
Aα − Sα = E(1) + µ E(2)
K α
− s2
− sα1 (µ − µα )
(K − K α )
102 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Hence
n
X
D̄ = 1+ fα (Aα − Sα )−1 : D
α=1
n
X fα
= E(1) + E(2) + E(1)
K
− sα1
α=1
(K − K α )
fα (2)
1 1 (2)
+ µ E : E(1) + E
− sα 3K 2µ
2
(µ − µα )
Finally,
n
1 X fα (1)
D̄ = 1 + E
3K K α
α=1 − s1
K − Kα
n
1 X fα (2)
+ 1 + µ E (6.42)
2µ − sα
α=1 2
µ − µα
n
( )
X
C= 1+ fα (Bα − Tα )−1 :C
α=1
where Bα is defined as
Bα = (D − Dα )−1 : D
1 (1)
Here D is the elastic compliance tensor of the matrix, i.e. D = 3K E +
1 (2)
2µ E . We can then calculate
Bα = (D − Dα )−1 : D
1 (2) −1
1 1 1 (1) 1 1
= − α E + − E :D
3 K K 2 µ µα
K − K (1) µα − µ (2) −1 1 (1)
α
1 (2)
= E + E : E + E
3KK α 2µµα 3K 2µ
Kα (1) µα
= − E − E(2)
K − Kα µ − µα
Subsequently,
Kα µα
Bα − T α = − − (1 − s α
1 ) E (1)
+ − − (1 − sα
2 ) E(2)
(K − K α ) (µ − µα )
K µ
α (1) α (2)
= − − s1 E − − s2 E
K − Kα µ − µα
and
−1 1 1
Bα − Tα =− E(1) − µ E(2)
K α
− s2
− sα1 (µ − µα )
(K − K α )
104 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Finally
n
X
C̄ = 1+ fα (Bα − Tα )−1 : C
α=1
n n
X fα X fα
= 1− E(1) + 1 − µ E(2)
K α
− s2
− sα1
α=1 α=1
(K − K α ) (µ − µα )
: (3KE(1) + 2µE(2) )
n n
X fα
(1)
X fα
= 3K 1 − E + 2µ 1 − µ E(2)
K α α
− s2
α=1 − s1 α=1
(µ − µα )
(K − K α )
Therefore
n
K̄ X fα
=1−
K K
α=1 − sα1
K − Kα
and
n
µ̄ X fα
=1− µ
µ − sα2
α=1
µ − µα
It is obviously that these results are different from the results obtained from
prescribed traction boundary condition. They are only agreeable to the first
order of the volume fraction. In other words, these two results (the results
obtained from prescribed stress b.c. and the results obtained from prescribed
strain b.c.) are not consistent in the homogenization scheme for dilute inhomo-
geneity distribution.
t = n · σ 0 , ∀ x ∈ ∂V
u(x) = x · 0 , x ∈ ∂V
0 =< > .
The second main difference between the self-consistent method and dilute
suspension method is that Eshelby’s equivalent inclusion principle is applied
with respect to the homogenized solid, instead of matrix. Suppose that there
are α = 1, 2, · · · , n distinct inhomogenous phases. ∀x ∈ Ωα ,
or
Dα : (σ 0 + σ d ) = C̄ : (σ 0 + σ d − σ ∗ ) (6.46)
Moreover, the disturbance field generated by eigenstrain is also calculated with
respect to homogenized solid, i.e.
d = S̄α : ∗ , ∀x ∈ Ωα (6.47)
σ d = T̄α : σ ∗ , ∀x ∈ Ωα (6.48)
106 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Therefore the average strain/stress inside the α-th phase inclusion may be
expressed by eigenstrain/eigenstress, i.e.
< >α = Āα : ∗ (6.49)
< σ >α = B̄α : σ ∗ (6.50)
where
Āα = (C̄ − Cα )−1 : C̄ (6.51)
B̄α = (B̄ − Bα )−1 : B̄ (6.52)
Subsequently, one can relate the average strain and average stress in the α-th
inclusion (inhomogeneity) with the background strain and background stress
by concentration tensors,
< >α = Āα : 0 (6.53)
< σ >α = B̄ α : σ 0 (6.54)
where the concentration tensors are defined as
Āα = Āα : (Āα − S̄α )−1 (6.55)
B̄ α = B̄α : (B̄α − T̄α )−1 (6.56)
Since by definition < σ >α = Cα :< >α and < >α = Dα :< σ >α , one
can rewrite Eqs. (6.53) and (6.54) as
α
C : Āα : D̄ : σ 0
< σ >α = (6.57)
B̄ α : σ 0
or α
D : B̄ α : C̄ : 0
< >α = (6.58)
Āα : 0
Note that the relationships 0 =< > and σ 0 =< σ > are used.
Suppose that prescribed macro-stress boundary condition is applied. Sub-
stituting Eqs. (6.57) and (6.58) into the basic average equation (6.11) yields,
α
Xn C : Āα : D̄ : σ 0
0 α
(D̄ − D) : σ = fα (D − D) : (6.59)
α
α=1 B̄ : σ 0
Therefore, self-consustent method gives the following estimate on effective
compliance tensor,
n
X
fα (Dα − D) : Cα : Āα : D̄
D +
α=1
D̄ = (6.60)
Xn
fα (Dα − D) : B̄ α
D+
α=1
Effective Elastic Modulus 107
D̄ : C̄ = 1 or D̄ = C̄−1 .
Consider
D = D : 1 = D : C̄ : C̄−1
h Xn i
= D: C+ fα (Cα − C) : Āα : C̄−1
α=1
n
X
= C̄−1 + fα D : (Cα − C) : Aα : C̄−1 (6.63)
α=1
Since,
D : (Cα − C) = D : Cα − 1
= −1 + D : Cα
= −(Dα − D) : Cα
which leads to
n
X
C̄−1 = C−1 + fα (Dα − D) : Cα : Āα : C̄−1 (6.65)
α=1
108 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Compare (6.65) with the first line of Eq. (6.60). One can conclude that
C̄−1 = D̄ (6.66)
Similar arguments can be made to show that D̄−1 = C̄.
Example 6.1 For isotropic composites, the effective moduli obtained from
self-consistent scheme can be further simplified.
Consider
Xn
C̄ = C + fα (Cα − C) : Āα (6.67)
α=1
Step 1.
which lead to
n Kα
K̄ X Kα −1
= 1+ −1 1+(
fα − 1)s̄1 (6.68)
K K K̄
α=1
n µα
µ̄ X µα −1
= 1+ fα −1 1+( − 1)s̄2 (6.69)
µ µ µ̄
α=1
3K − 2µ
Note that ν = .
2(3K + µ)
Lemma 6.2 (Tanaka and Mori) Consider two coaxial, similar ellipsoidal
domains, Ω0 , Ω (Ω0 ⊂ Ω),
x21 x22 x23
Ω0 = x + + ≤ 1
a21 a22 a23
x21 x22 x23
Ω = x 2 + 2 + 2 ≤1 (6.70)
b1 b2 b3
where
a1 a2 a3
+ + =k
b1 b2 b3
Assume that a uniform eigenstrain state, ∗ij (x), is prescribed in the smaller
ellipsoidal region, i.e.
∗
∗ ij x ∈ Ω0
ij (x) =
0 x ∈ IR3 /Ω0
The the average disturbance strain field is zero, i.e
Z
1
< >Ω−Ω0 = ij (x)dΩ = 0 . (6.71)
Ω − Ω0 Ω−Ω0
Proof:
Suppose that there are three coaixial, similar ellipsoidals, Ω0 ⊂ Ω1 ⊂ Ω2 in
an infinite homogeneous medium, and a uniform eigenstrain is presecibed in
Ω0 , i.e. ∗
∗ ij x ∈ Ω0
ij (x) =
0 x ∈ IR3 /Ω0
The disturbance displacement field can be then written as
Z
ui (x) = − ∗mn Ck`mn Gik,` (x − x0 )dx0 (6.72)
Ω0
Since Eshelby tensor only depends on the material property and the aspect ratio
of the ellipsoidals,
h i
∗mn Ω0 Sk`mn
Ω2 Ω1
− Sk`mn =0 (6.76)
112 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
♣
Remark 6.3.1 1 It is also true that the average disturbance stress field is
also zero Z
σij dΩx = 0 . (6.79)
Ω−Ω0
let to
0 + d = A Ω : ∗ , ∀ x ∈ Ω (6.85)
where AΩ = (C − CΩ )−1 : C. Combining with d = SΩ : ∗ , one can find
that
∗ = (AΩ − SΩ )−1 : 0 (6.86)
Substitute (6.86) back to (6.83). We finally have
< >Ω = 1(4s) + SΩ : (AΩ − SΩ )−1 : 0 (6.87)
The average stress inside the inclusion can be also evaluated by considering
homogenization condition and (6.86)
< σ >Ω = C : 0 + d − ∗ = C : 0 + (SΩ − 1(4s) )∗
= C : 1(4s) + (SΩ − 1(4s) (AΩ − SΩ )−1 : 0 . (6.88)
One the other hand, by the Tanaka-Mori lemma, the average strain in the
matrix is
< >M =< 0 + d >M = 0 (6.89)
114 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
and hence
< σ >M = C : 0 . (6.90)
Let f be the volume fraction of the inhomogeneity. We then have the following
balance equations for average strain and stress
and
< σ >V = (1 − f )C : 0 + f C : 1(4s) + (SΩ − 1(4s) (AΩ − SΩ )−1 : 0 .
= C : 1(4s) + f (SΩ − 1(4s) (AΩ − SΩ )−1 : 0 . (6.94)
By definition,
< σ >V = C :< >V (6.95)
It leads to
C : 1(4s) +f (SΩ −1(4s) )(AΩ −SΩ )−1 : 0 = C̄ : 1(4s) +f SΩ (AΩ −SΩ )−1 : 0
Let pt and σ pt representing perturbed strain and stress fields due to Es-
helby’s single inclusion solution in an infinite medium. If we let
Obviously,
σ 0 + σ pt 6= σ 0 , or 0 + pt 6= 0 , ∀ x ∈ ∂V (6.101)
Therefore, either boundary condition (6.98) and (6.97) will not be satisfied.
This is because a finite size RVE will cause additional interaction between
matrix and inclusions, interaction between the boudary and inclusions, and
interaction among inclusions themself. Note that pt , σ pt , → 0 only when
|x| → ∞.
To take into account the effects of a finite size RVE, additional stress and
strain fields are need to faithfully represent total stress and strain distribution
in an RVE, i.e.
σ = σ 0 + σ̃ + σ pt (6.102)
= 0 + ˜ + pt (6.103)
where σ̃ and ˜ are the so-called image stress and image strain.
In literature, especially literatures on dislocations, additional stress and strain
fields that accommodate the stress solution of a infinite space to satisfy bound-
ary conditions are called image stress and image strain fields, because in prac-
tice some of these stress and strain fields are found by placing certain image
external sources to achieve their objectives.
Nevertheless, the homogenization problem in a finite REV becomes com-
plicated, because in general it is very difficult to know the precise distribution
of image stress and image strain fields. To circumvent this difficulty, Mori
and Tanaka [1973] proposed the following mean field assumption, which is an
ingenous and very successful method.
Mori & Tanaka’s theory was later refined in a landmark paper by G. Weng
(Weng [1990]). The following presentation is an adaption of Weng’s formula-
tion. Suppose that in an RVE there are many inhomogeneities, or the density
of inhomogeneities are statistically stable. Then the strain or stress field in the
matrix may be written as
In general we don’t know the precise disturbance fields in a matrix, i.e., dM or
σ dM .
116 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
and
−1 −1
h i
dil
BΩ = BΩ : (BΩ − TΩ )−1 = (BΩ − TΩ ) : AΩ
−1 −1
h i
= 1 − TΩ : BΩ
h i−1
= 1 − TΩ : D−1 : (D − DΩ )
h i−1
= 1 + QΩ : (DΩ − D) (6.112)
where
PΩ = SΩ : C−1 (6.113)
QΩ = TΩ : D−1 (6.114)
By definition,
where
h i−1
Ã0 := fM Adil M + fΩ Adil Ω (6.120)
h i−1
B̃0 := fM B dil M + fΩ B dil Ω (6.121)
Accordingly,
< >Ω = Adil Ω :< >M = Adil Ω : Ã0 :< > (6.122)
< σ >Ω = B dil Ω :< σ >M = B dil Ω : B̃0 :< σ > (6.123)
Therefore,
and
(a) (b)
(c) (d)
(e) (f)
Figure 6.4. Comaprison of effective bulk modulus among various homogenization methods:
dilute distribution (DD & DT), self-consistent, and Mori-Tanaka
120 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
In general, for a solid with n+1 phases (from α = 0 to α = 0), the Mori-Tanaka
mean field theory gives the following estimates,
n
X n
X −1
C̄ = fα Cα : Adil α : fα Adil α
α=0 α=0
n
X n
X −1
D̄ = fα Dα : B dil α : fα B dil α (6.126)
α=0 α=0
where the pahse α = 0 represents the matrix, and non-zero α represents the
inhomogeneous phases.
Rodney’s mother had been a student at Leeds School of Art. Rodney himself
was also an only child, in an immediate home background which encouraged
scholarship and self-sufficiency.
Rodney entered Leeds Grammar School with a scholarship in 1932, and
there gave regular prize-wining evidence of all-round intellectual ability not
only in mathematics, but equally in art, English literature, and other Arts sub-
jects. During this period he taught himeself to play the piano, and became
proficient at chess in which he was later to represent Cambridge University
and town. Thus were developing the powers of accurate observation and anal-
ysis to be brought to bear on the mathematics and physics which became his
formal specialism from the age of 15. The customary large-team games did
not attract him as school, but Rodney enjoyed the one-to-one sports of squash,
fencing, and golf. He left school as Head of House, and in December 1938
he was awarded an Open Major Scholarship at Pembroke College, Cambridge.
However, it needed the State and County Scholarships gained in the preceding
summer to make a financially independent undergraduate.
Hill went up to Cambridge to read Mathematics in October 1939, againt a
background of external events which must have seemed the least auspicious
since the very founding of the University. Major Scholars were expected to
take Part II of the Tripos in two years instead of three by omitting all first-year
courses. This imposed a heavy workload, to be carried under spartan condi-
tions created by wartime restrictions such as blackout and rationing combined
with antique College plumbing. For example, there was no running hot wa-
ter, the nearest bath was courts away, and the winter allocation of one sack of
coal per week fuelled a fire in one’s room only in the evenings. Hill was not
deflected by the adverse general situation from his aim of a first-class honours
degree, and he became a Wrangler in June 1941. This entitled him to take
Part III of the mathematical Tripos, in the applied mathematical part of which
quantum mechanics figured prominently at the time. However, he felt obliged
to war-work, and so lost the opportunity for advanced training which those
lecture courses would have provided.
........
Problems brought to the Theoretical Research Branch were distributed ini-
tially according to specialisms of the more senior members, some of whom
had acquired relevant experience at Woolwich Arsenal. Those problems which
were quite new in context tended to go to the young inexperienced graduates
newly arrived from university. This was indeed a baptism of fire for them,
but it was a test which was to reveal Hill’s true metier. One of his initial as-
sigments was the deep penetration of very thick armour by Munoroe jets and
high-velocity shells with tungsten-carbide cores. This required a mechanics
of plastic deformation with unlimited magnitude, and thus was aroused Hill’s
interest in the field in which he later became perhapse the foremost world au-
122 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
thority. At this stage, however, he had no prior knowledge of the physics and
metallurgy of plasticity, and little of stress, strain or the tensors which the
mathematics would eventually require. There was no useful textbook, but G.
I. Taylor had written one or two helpful reports on shaped charges and Munros
jets. Nevertheless, working at first with Mott and Pack, Hill was soon able
to show, for example, that penetration by a tungsten–carbide core with pure
ogival head would be seriously degraded if too much of the tip were ground
conical (the British practice for manufacturing convenience). The demonstra-
tion was achieved not only theoretically, but also in field trials planned by Hill
in collaboration with an experimental group under Dr. Charles Sykes, F.R.S.
The problems at Fort Halstead called for simple but effective mathemat-
ics guided by physical intuiation and a willingness to communicate with oth-
ers, including non-mathematicians and experimentalists. There was not time
for complicated mathematics, there were no electronic computers to assist it,
and the experimental data were ususally too crude to warrant it anyway. He
acquired a lasting taste for a pragmatic blend of rigour, elgance, and simple
realism in the application of mathematics.
The sense of purpose discovered at this time was noticed by colleagues as
a cheerful and sparking earnestness. Popular relaxations among the group at
Cambridge had included music, books, and lightning chess. At Fort Halstead
ballroom dancing was added as a consuming passion for some, and Hill was
not slow to find that he had medal-winning ability in this new enthusiasm. He
met his future wife, Jeanne Wickens, early in 1945. She had been transferred
to work at Fort Halstead from the bombing range at Shoeburyness. Previously
she had trained as a dancer and teacher of ballet, but war cut short a promising
career. They were married in Cambridge in 1946, and they have one duaghter,
born in 1955. The strength of his wife’s support could already be detected in
the Preface to Hill’s first book, and the passage of years has happily reinforced
this bond.
By this time the applied mechanics of both solid and fluids was being forced
to push the boat out onto a sea of nonlinear problems, and away from the haven
linearity in which much pre-war work had lingerd. The trend was evident not
only in England, of course, but in other countries too. Hill found himself in
demand as the sole adviser on continuum plasticity in England, not only con-
cerning problems arising from the interests at Fort Halstead, but also fot new
theories of metal-working processes needed by engineers in the steel indus-
try. He obtained a Cambridge Ph.D. in 1948 for a Thesis entitled “Theoretical
studies of the plastic deformation of metals”. From the Ph.D. Thesis grew a
much more extensive monograph on “The Mathematical Theory of Plasticity”,
published at the Clarendon Press, Oxford, in 1950. This very rapidly estab-
lished Hill as an international authority. The final draft was written in his spare
time, i.e. in the evenings and weekends. He was then still only in his 28th year,
Effective Elastic Modulus 123
and it is timely to recall a remark from the review of the book in Engineering:
“The author has done his work so well that it is difficult to see how it could be
bettered. The book should rank for many years as an authoritative source of
reference.” This prognostication was fully borne out. The book was in print at
Oxford for 21 years, Japanese and Russian translations have been made, and
total sales currently approach 13,000.
The Journal of Mechanics and Physics of Solids was launched with the en-
couragement of the infan Pergamon Press in 1952. Hill suggested the title and
the general aim of a forum for effective applied mathematics, linked with ex-
perimenation, in engineering science. From the onwards the Journal has been
regarded as among the foremost in its general field, and unique in flavor. Hill
served as Eidtor-in-Chief untill handing over in 1968 to H.G. Hopkins.
The University of Nottinggham had received its Chater, and independence
from London, in 1948, and was shortly to embark on two decades substantial
expansion. Professor H. R. Pitt was appointed in 1950 to head the existing
Mathematics Department, and he was soon instrumental in securing the cre-
ation of a new Chair of Applied Mathematics. Rodney Hill applied, and was
offered the post in 1953 while still on 31. It was his responsibility to modernize
the teaching of applied mathematics. Hill took over some existing course him-
self, and instigated new ones with the aim of encouraging research students.
His undergraduate lectures were characterized by conciseness and tendency to
brevity. He would never exceed the time limit. But those stidents who took
the trouble to write down what he said, in addition to what was written on the
blackboard, found after reflection that they had a first-calss and substantial set
of notes.
It may only have been a coincidence that emergence of interest in the so-
called rational continuum mechanics was taking place in some American and
British universities at this time. Hill’s writings demonstrate an independent
view of these development, and no taste at all for axiomatics. He was beginning
to lay down the basis of general studies of non-uniqueness and instability in
continua which were to prov highly influential over the next two decades, and
which in due course brought further students and able collaborators.
The University of Cambridge conferred the degree of Sc. D. upon Rodney
Hill in 1959. The highest honour to which any British scientist aspires followed
in 1961, when he was elected a Fellow of the Royal Society. This gave much
pleasure to his colleagues at Nottingham and to his friends elsewhere.
In 1963 Hill was elected to a Berkeley Bye-Fellowship at Gonville and Caius
College, Cambridge. This he held for 6 years until the University conferred a
personal Readership in Mechanics of Solids. Thus he became a member of
the teaching staff of the Department of Applied Mathematics and Theoretical
Physics, and in 1972 a personal Professorship was conferred.
124 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
6.5 Exercises
Probelm 6.1 Consider a n-phase composite material, and each phase has
its own elastic tensor Cα , compliance tensor Dα ; and matrix has elastic ten-
sor, C, and compliance tensor, D. Assume that in the representative volume
element (RVE), each phase only appears as one ellipsoidal inclusion. Under
dilute distribution assumption, the corresponding Eshelby tensor and conju-
gate Eshelby tensor for each phase are Sα and Tα respectively. Denote
Aα = (C − Cα )−1 : C (6.127)
Bα = (D − Dα )−1 : D (6.128)
Show
Cα : Aα : (Aα − Sα )−1 : D = Bα : (Bα − Tα )−1 (6.129)
Dα : Bα : (Bα − Tα )−1 : C = Aα : (Aα − Sα )−1 (6.130)
Probelm 6.2 For an isotropic two phase material. Assume the inhomogene-
ity phase is random distributed spherical cavities (µI = 0; KI = 0), and
the matrix is an incompressible masterial (K → ∞). Use the self-consistent
scheme,
n Kα
K̄ X Kα −1
= 1+ fα −1 − 1)s̄11+( (6.131)
K K K̄
α=1
n µα
µ̄ X µα −1
= 1+ fα −1 1+( − 1)s̄2 (6.132)
µ µ µ̄
α=1
Effective Elastic Modulus 125
where
1 + ν̄
s̄1 = (6.133)
3(1 − ν̄)
2(4 − 5ν̄)
s̄2 = (6.134)
15(1 − ν̄)
to find the effective bulk modulus, K̄, and the effective shear modulus, µ̄.
Hint:
J.R. Willis, “Variational and related methods for the overall properties of
composite”, in Advance in Applied Mechanics, Edited by C.-S. Yih (pages 45-
46), (1981), Academic Press, New York.
B. Budiansky, “On the elastic moduli of some heterogeneous materials”,
Journal of Mechanics and Physics of Solids, Vol. 13, (1965), pages 223-227.
Probelm 6.3 Assume that in an RVE there are n+1 phases, α = 0, 1, · · · , n
Mori-Tanaka mean theory states that
X n n
X −1
dil
D̄ = fα Dα : B α : fα B dil α (6.135)
α=0 α=0
Xn n
X −1
C̄ = fα Cα : Adil α : fα Adil α (6.136)
α=0 α=0
Show that Mori-Tanaka scheme is self-consistent, i.e.
C̄ = D̄−1 (6.137)
Hint: First show that
Cα : Adil α = B dil α : C0 (6.138)
Dα : B dil α = Adil α : D0 (6.139)
Probelm 6.4 Consider a two-phase composite with randomly distributed
spherical inclusions. The ratios of material constants between inhomogeneity
and matrix are
KΩ
= 25, and K Ω = 750M Pa (6.140)
K
νΩ
= 4, and ν Ω = 0.4 (6.141)
ν
K̄ µ̄ ν̄
Plot the ratio of , , and verses the volume fraction of inhomogeneity,
K µ ν
f , by using homogenization methods under the assumption of dilute suspen-
sion (both prescribed traction and prescribed displacement), self-consistent
method, and Mori-Tanaka mean field method.
126 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Chapter 7
Figure 7.1. Illustraions of dislocations: (a)edge dislocation, and (b) screw dislocation
That is the traction is required to be continuous across the slip plane. However,
the solution of such boundary-value problem is difficult, and people have not
found any important applications of such dislocation model.
w(r, θ) = Bθ (7.22)
Assume that the length of the hollow cylinder is L. The energy per unit
length in z-direction is,
2
1 L 2π R σzθ 2
Z Z Z Z
W 1 σzθ
= dV = rdrdθz.
L L V 2µ V 0 0 r0 2µ
b2 µ R d b2 µ R
Z
= = ln . (7.32)
4π r0 r 4π r0
First, as R → ∞, W/L → ∞. This shows that the self-energy of the dis-
location depends on the size of the crystal. On the other hand, for a finite
size crystal, the dislocation solution of unbounded domain does not hold true
because the image stress caused by the boundary.
Assume that the dislocation is far away from the boundary, the boundary
effecrts are abated inside, one may choose the dimension of the crystal, say `
as R; in polycrystallines, one may choose the size of a grain as R, where the
dislocation resides.
Second, as r0 → 0, W/L → −∞. This abnormality is due to the limitation
of linear elasticity model. Within five atomic spacing of a dislocation core, the
linear elasticity model is no longer valid. In general, the length of the Bergurs
vector is close to the lattice spacing. Therefore, in practice, we usually choose
r0 = 5b or r0 = b/α, 0 < α < 1 such that the elastic self-energy equals to
W µb2 ` W µb2 α`
= ln , or = ln . (7.33)
L 4π 5b L 4π b
By defnition, the self-energy should include the core energy, i.e.
W self = W elas + W core (7.34)
The core energy is relatively small, but may not be negligible, because it is
10% to 20 % of the elastic self-energy. It may be relatively small, but can not
be neglected. Overall, the linear elasticity theory gives a good estimate of self-
energy. In Sec. 4 of this Chapter, we shall discuss the Peierls-Nabarro model,
which provides a means to estimate the core energy.
∞ bµ y
σxz (x, y) = − (7.35)
2π (x + `)2 + y 2
∞ bµ (x + `)
σyz (x, y) = . (7.36)
2π (x + `)2 + y 2
132 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
I bµ y
σxz (x, y) = (7.38)
2π (x − `)2 + y 2
I bµ (x − `)
σyz (x, y) = − . (7.39)
2π (x − `)2 + y 2
The total stress distribution is then the superposition of the solution in the infi-
t = σ∞ + σI ,
nite space and the solution of of image stress distribution, i.e. σij ij ij
where the superscript, t, ∞, and I denote the total stress solution, the solution
obtained in the infinite space, and the image stress solution.
By anti-symmetry, the traction-free boundary condition at x = 0 is then
enfored,
t ∞ I by y by y
σxz (0, y) = σxz (0, y)+σxz (0, y) = − + ≡ 0 . (7.40)
2π `2 + y 2 2π `2 + y 2
Introduction of Dislocation Theory 133
Remark 7.1.1 1. Note that the image stresses at x = −` and y = 0, i.e. the
position of the real dislocation, are
I I bµ
σxz (−`, 0) = 0, σyz (−`, 0) = . (7.41)
4π`
2. When |x|, |y| >> `,
t t
σxz (x, y) ≈ 0, and σyz (x, y) ≈ 0, (7.42)
which means that outside the region of {(x, y) (x + `)2 + y 2 ≤ `2 }, the total
stress is almost negligible.
∇2 ∇2 ψ = 0 . (7.50)
Because the defect configuration, for an edge dislocation, the region right
above around the dislocation core should be in compression, whereas the re-
gion right below the dislocation core should be in tension, i.e.
Then,
∂2 1 ∂ 1 ∂2
+ + ψ = β1 r−1 sin θ . (7.55)
∂r2 r ∂r r2 ∂θ2
Introduction of Dislocation Theory 135
b h −1 y λ+µ xy i
u(x, y) = − tan + (7.69)
2π x λ + 2µ x2 + y 2
b h µ λ+µ y2 i
v(x, y) = − − ln(x2 + y 2 ) + (7.70)
2π 2(λ + 2µ) λ + 2µ x2 + y 2
∞ µb y(3(x + `)2 + y 2
σxx = − (7.71)
2π(1 − ν) ((x + `)2 + y 2 )2
∞ µb y((x + `)2 − y 2
σyy = (7.72)
2π(1 − ν) ((x + `)2 + y 2 )2
∞ µb (x + `)((x + `)2 − y 2
σxy = (7.73)
2π(1 − ν) ((x + `)2 + y 2 )2
Introduction of Dislocation Theory 137
will not satisfy the traction-free boundary condition at x = 0 i.e. σxx (0, y) 6= 0
and σxy (0, y) 6= 0.
If we place a fictitous dislocation at x = ` with the opposite Burgers vector.
The induced image stress fields,
I µb y(3(x − `)2 + y 2
σxx = (7.74)
2π(1 − ν) ((x − `)2 + y 2 )2
I µb y((x − `)2 − y 2
σyy = (7.75)
2π(1 − ν) ((x − `)2 + y 2 )2
I µb (x − `)((x − `)2 − y 2
σxy = (7.76)
2π(1 − ν) ((x − `)2 + y 2 )2
∞ (0, y)+σ I (0, y) =
will cancel the normal stress on traction-free surface, i.e. σxx xx
0, but it can not cancel the shear stress at x = 0. In fact,
∞ I µb `(`2 − y 2 )
σxy (0, y) + σxy (0, y) = 6= 0 . (7.77)
π(1 − ν) (`2 + y 2 )2
To cancel the shear stress on traction-free surface, one has to superpose another
stress field, such that the third stress fields satisfy the condition,
d4 Ψ̄s 2
2 d Ψ̄s
− 2ξ + ξ 4 Ψ̄s = 0. (7.82)
dx4 dx2
Solving (7.82) yields the following solution,
Ψ̄s (x, ξ) = (a0 (ξ) + a1 (ξ)) exp(ξx) + (b0 (ξ) + b1 (ξ)) exp(−ξx) . (7.83)
138 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
1 ∞
Z
Ψ(x, y) = a1 (ξ)x exp(ξx) sin(ξy)dξ . (7.86)
π ∞
and the definition of the Fourier-cosin transform, one may find that
∞
`(`2 − y 2 )
Z
1 µb
a1 (ξ)ξ = 2 2 2
cos(ξy)dy
π −∞ π(1 − ν) (` + y )
Z ∞
µb `(`2 − y 2 )
= exp(iξy)dy . (7.88)
π 2 (1 − ν) −∞ (`2 + y 2 )2
∞
`(`2 − y 2 )
Z
The last line is because of 2 2 2
sin(ξy)dy = 0.
−∞ (` + y )
Use the residue theorem to evaluate the integra,
∞
`(`2 − y 2 )
Z X
2 2 2
exp(iξy)dy = 2πi Res F (yN )
−∞ (` + y ) yN =i`
iξ`
= 2πi − exp(−ξ`) = πξ`exp(−ξ`) . (7.89)
2
Wer then find that
µb`
a1 (ξ) = exp(−ξ`) (7.90)
π(1 − ν)
so that
Z ∞
µb`
Ψ(x, y) = x exp ξ(x − `) sin ξydξ
π(1 − ν) 0
µb`xy
= (7.91)
π(1 − ν)[(x − `)2 + y 2 ]
Introduction of Dislocation Theory 139
and
000 µb` (`2 − x2 )y 2 y 2 (3x2 − (y + `)2
σxy = − + (7.92)
π(1 − ν) [(x − `)2 + y 2 ]2 [(x − `)2 + y 2 ]3
000 2µb`xy
σxx = − [3(` − x)2 − y 2 ] (7.93)
π(1 − ν)r6
I 000 µb
σxy (−`, 0) + σxy (−`, 0) = . (7.95)
4π(1 − ν)
dA = dX × δη . (7.96)
All the atoms on this area will be sujected a discontinuous jump with the
direction and the magnitude of the local Burgers vector, b. The traction forces
on the infinitesimal area can be expressed as σ · dA. Be precise, it is
σ · dA = σ · (dX × δη) (7.97)
If we assume that the work done by the stresses relates to the decreases of the
potential energy of the dislocation,
d(δE) = −b · σ · (dX × δη) (7.98)
The change of the total energy due to the virtual displacement field is
Z Z
δE = − b · σ · (dX × δη) = − (σ · b) × td` · δη (7.99)
L L
where dX = td`.
By definition, the decrease of the potential energy under the virtual displace-
ment field is the external virtual work done along the dislocation loop, i.e.
Z
δE = −F · η = − F` d` · δη , (7.100)
L
where F` is the force per unit length along the dislocation loop.
Hence, we derived the celebrated Peach-Koehle equation,
Z
F= σ · b × td`, and F` = σ · b × t . (7.101)
L
where F` is the force per unit length. In the case of straight dislocation line,
F
we often denote it as .
L
Now, let’s look at a few examples.
To simplify the computation, we denote
g := σ · b. (7.102)
Then the Peach-Koehle force formula can be conveniently written into a matrix
form,
e1 e2 e3
F` = g × t = g1 g2 g3 . (7.103)
t1 t2 t3
Example 7.1 This example is illustrated in Fig. 7.5. We are examing the
external forces exerted on a straight screw dislocation.
Let x = 1, y = 2, z = 3. In this case, the unit vector of the dislocation line
is t = ez , the Burgers vector is b = bez , and the stresses other than self-stress
are
σ = σxz ex ⊗ ez + σzx ez ⊗ ex + σyz ey ⊗ ez + σzy ez ⊗ ey . (7.104)
Introduction of Dislocation Theory 141
and
gx = σxz b, gy = σyz b, gz = 0 . (7.105)
Hence
ex ey ez
F` = g × u = σxz b σyz b 0 = σyz bex − σxz bey . (7.106)
0 0 1
To interprete the meanings of this expression, we would say that the shear
stress, σxy , moves the dislocation line to +x direction, whereas shear stress,
σxz moves the dislocation line towards the negative direction of Y-axis, i.e. -Y
direction.
This is to say that the shear stress, σxy , will move the dislocation line along
the slip plane in the positive direction of X-axis. On the other hand, the normal
stress, σxx , will make the dislocation line tranlating along its own direction.
This is an unconservative motion, because if the motion is addmissible, one
has to remove material at one end of dislocation line and add material (atoms)
at the other end of the dislocation line. In literature, we refer such dislocation
movement as “climbing”.
From Eq. (7.109), one may find that if σxx < 0, which means the material is
under compression, the Peach-Koehle force will squeeze the dislocation line up
in Y-axis, and when σxx > 0 it will pull the material apart and let dislocation
line climbing down.
Introduction of Dislocation Theory 143
II µb2 (y − y0 )
σxz = − , (7.111)
2π (x − x0 )2 + (y − y0 )2
II µb2 (x − x0 )
σyz = . (7.112)
2π (x − x0 )2 + (y − y0 )2
In this case, the Peach-Koehle force equation is
F` = σyz ex − σxz ey . (7.113)
1. Calculate the force, F1→2
` , which is the force exeretd on the dislocation, S2 ,
by the dislocation, S1 . Let r = r0 and θ = θ0 in (7.110) and substitute them
into (7.113). We have
F1→2
`
I
= σyz I
b2 ex − σxz b 2 ey
x0 ,y0 x0 ,y0
µb1 b2 cos θ0 µb1 b2 sin θ0
= ex + ey
2π r0 2π r0
µb1 b2 µb b
1 2
= cos θ0 ex + sin θ0 ey = r̄0 , (7.114)
2πr0 2πr0
where r̄0 = r0 /|r0 | is the unit vector in r0 direction.
2. Calculate the force exerted on the dislocation S − 1 by the dislocation
S2 . In this case, we let x = 0, y = 0 in (7.111) and (7.112) and substitute them
into (7.113),
F2→1
`
II
= σyz II
b1 ex − σxz b 1 ey
0,0 0,0
µb1 b2 cos θ0 µb1 b2 sin θ0
= − ex − ey
2π r0 2π r0
µb1 b2 µb1 b2
= − cos θ0 ex + sin θ0 ey = − r̄0 . (7.115)
2πr0 2πr0
It is obvious that F1→2
` = −F2→1
` (see Fig. 7.7).
We then conclude that when b1 and b2 are along the same direction, the two
screw dislocation repel each other, if b1 b2 < 0, i.e. b1 and b2 are in opposite
direction, then the two screw dislocations attract to each other.
144 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
1 Do not confusion this with the statically admissible virtual forces in continuum mechanics.
Introduction of Dislocation Theory 145
From Fig. 7.8, one may observe that the area difference between V 0 and V is
ω1 − ω2 , i.e. adding the area ω1 and removing the area ω2 . Hence the stored
strain energy difference is
Z Z
δE1 = W dV − W dV . (7.121)
ω1 ω2
Since δX is infinitesimal,
Z Z
ω1 − ω2 = dA = −δX · dsn
ω1 −ω2 L
Note that in this step, all the operations are performed in the material config-
uration. We are comparing the energy difference between two adjacent local
material volumes differing a translation.
(3) During a configuration change, the defect moves +δX from its original
material position to the new material position, it will cause the relative material
virtual displacement,
∂ui
δui = δXj . (7.123)
∂Xj
Introduction of Dislocation Theory 147
0
This is to say that if there is no displacements along ∂V , the displacement on
∂ui
∂V is δui = δXj ∀X ∈ L. Then the difference of the work done to the
∂Xj
environment of the two local configurations is:
Z Z Z
ext
δW = 0 · Ti ds − δui Ti ds = − δui σij nj ds
0
LZ L L
PE = W 1(2) − E ⊗ D (7.129)
PM = W 1(2) − ∇u ⊗ σ . (7.130)
and
∂W
6= σmn um,nk .
∂xk
Suppose that there is a defect at a material point ξi , we assume that this may
be captured by an equivalent inhomogeneous elastic stiffness tensor Cijk` (X−
ξ), i.e. 0
Cijk` , ∀X 6= ξ
Cijk` (X − ξ) = (7.133)
Cijk` (ξ), ∀X = ξ
where
∂2W ∗ 1
0
Cijk` (ξ) = Cijk` − and W ∗ = C ijk` ∗ij ∗k` (7.134)
∂ij ∂k` 2
Consider Cijk` ui,j = σk` and integration by parts for the second term of the
integrand.
Z h
inh 1 i
Fn = Cijk` ij k` − (σk` uk,n )` + σk,`` uk,n dV
2 ,n
ZV h
1 i
= Cijk` ij k` − (σk` uk,n ),` dV
2 ,n
IV I
= W δn` − uk,n σk` n` ds = Pn` n` ds . (7.136)
L L
Example 7.4 The asymptotic stress fields for a mode III crack is
KIII θ KIII θ
σ13 = − √ sin , σ23 = √ cos . (7.137)
2πr 2 2πr 2
150 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
if Ω = IR3 , we have
Z ∞ Z ∞ Z
0 0 0 0 0
χ(x )δ(s − n · x )dx = δ(s − n · x )dx = dS (7.142)
−∞ −∞ S
∗
ui,j = βij + βij (7.149)
Introduction of Dislocation Theory 153
The above expression was derived by Volterra, and it is called Volterra formula
(Volterra [1907]).
Differentiating (7.156) yields
Z
ui,j (x) = Cj`mn bm nn Gij,`j (x − y)dSy (7.157)
S
2 There are many attempts to derive plasticity theory from this formualtion.
154 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Mura showed (Mura [1963]) that the above surface integration can be written
as a line integration,
I
βji (x) = ejnh Cpqmn Gip,q (x − y)bm th d`y (7.159)
L
or in component form
Z I
ek`h nk Ajih,` dS = th Ajih d` (7.161)
S
∂
where Gip,q` = − Gip,q . Utilizing the identity ek`h ejnh = δkj δ`n − δkn δ`j ,
∂x0`
one can obtaion
Z
− (δkj δ`n − δkn δ`j )nk bm Cpqmn Gip,q` (x − x0 )dS 0
ZS
= − nj bm Cpqm` Gip,q` (x − x0 ) − nn bm Cpqmn Gip,qj (x − x0 ) dS
Z S
= nj bm δim δ(x − x0 ) + nn bm Cpqmn Gip,qj (x − x0 ) dS 0
ZS Z
= nj bi δ(S − x0 )δ(x − x0 )dx0 + nn bm Cpqmn Gip,qj (x − x0 )dS 0
Ω S
Z
= nj bi δ(S − x) + nn bm Cpqmn Gip,qj (x − x0 )dS 0 (7.163)
S
Let,
∂
∇= em , A = A, ··· en , dS = dSk ek , and d` = tk d`ek = dxk ek .
∂xm
A special case of the Stoke’s theorem is,
Z I
∂A,···
mnk dSk = A,··· dxn . (7.171)
S ∂xm ∂S
We then have
Z I
−ijk mnk A,···m dSn = ijk A,··· dxk
S ∂S
Z I
−(δim δjn − δin δjm ) A,···m dSn == ijk A,··· dxk (7.173)
S ∂S
In (7.169), we may view R,pp as A,pp in the second integral and R,mp as
A,mp in the third integral and then apply the Stoke’s theorem (7.174) to (7.169),
Z Z
0 0 0 0
b` R,pp` dSm − R,ppm dS` = −b` R,pp`0 dSm − R,ppm0 dS`
S S
I
0
= −b` m`k R,pp dxk
C
Z Z
0 0 0 0
bj R,pmp dSj − R,pmj dSp = −bj R,pmp0 dSj − R,pmj 0 dSp
SI S
0
= −bj jpk R,pm dxk
C
λ+µ 1
In the last line, the identity = is used. Consider the fact that
λ + 2µ 2(1 − ν)
0
xj − xj Rj δmp Rm Rp
R,j = = , and R,mp = −
R R R R3
hence
2 −2Rj
R,pp = and R,ppj = .
R R3
Therefore,
Z I
1 bm R j 0 1 m`k b` 0
um (x) = − 3
dSj − dxk
4π S R 4π C R
I
1 ∂ Rp 0
− pjk bj dxk (7.176)
8π(1 − ν) C ∂xm R
which can be put into an elementary vector form, i.e. the Burgers formula
0 0
b × d` b × R · d`
Z I
b 1 1
u(x) = − Ω − − ∇ . (7.177)
4π 4π C R 8π(1 − ν) C R
0 0
In (7.177), d` = tk d`ek = dxk ek , and Ω is the so-called solid angle, which
is defined as the surface area Ω of a unit sphere covered by the surface’s pro-
jection onto the sphere. In this case, the angle is subtended by the dislocation
surface, S, i.e.
0 0
Rj dSj n · dS
Z Z
Ω= = (7.178)
S R3 S R2
where n := R/R is a unit vector from the point x to the dislocation surface,
S.
If the surface is a sphere, dS = R2 dω and
R2 n · dω
I I
Ω = = n · dω
S2 R3 S2
I
= ni ni dω = 4π . (7.179)
S2
In the above equation, only the first term is not a line integral. Nevertheless,
we claim that
Z I
0 0
bm R,pp`j dSj = −8πδ(S − x)bm n` − bm j`k R,ppj dxk .
S C
Proof:
Apply Stokes’ theorem,
I Z h i
0 0
ijk φdxk = φ,j dSi − φ,i dSj (7.181)
C S
to the above expression,
I Z
0 0 0
i`k R,pp dxk = R,pp`0 dSj − R,ppj 0 dS`
C ZS
0 0
= R,ppj dS` − R,pp` dSj (7.182)
S
Therefore,
I Z h
∂ 0 0 0
i
j`k R,pp dxk = R,ppjj dS` − R,pp`j dSj (7.183)
∂xj C S
Since
0 1 0
GP (x − x ) = , and ∇2 Gp = −δ(x − x ),
4πR
we then have
2 0 0 0
R,pp = = 8πGP (x−x ) and R,ppjj = 8π∇2 GP (x−x ) = −8πδ(x−x ).
R
Consequently,
I Z Z
0 0 0 0
bm j`k R,ppj dxk = −8πbm δ(x − x )dS` − bm R,pp`j dSj
C S S
Use Radon transformation,
Z Z
0 0 0
δ(x − x )dS` = δ(x − x )n` dS
ZS S
0 0 0
= δ(x − x )n ` δ(S − x )dΩ = δ(S − x)n` (7.184)
IR3
Hence, we verfied the claim.
∗ = −8πb n δ(S − x), we again recover Mura’s formula
Note that βm` m `
I
∗ 1 0
βm` = um,` − βm` =− j`k bm R,ppj dxk
8π C
I I
1 0 1 0
− mnk bn R,pp` dxk − jpk bj R,mp` dxk (7.185)
8π C 8π(1 − ν) C
Introduction of Dislocation Theory 159
where b is the Burgers vector, t is the tangential vector along the dislocation
loop, and
I e
Σij = σij + σij (7.195)
I are the stress fields of all other dislocation loops inside the solid,
Here σij
which can be expressed as
I
I µ h1 0 0
σij = bn R,mpp (jmn dxi + imn dxj )
4π C 2
1 i 0
+ kmn (R,ijm − δij R,ppm ) dxk (7.196)
1−ν
Introduction of Dislocation Theory 161
e is the stress field due to externally applied loads.
and σij
Denote
fiP K = ijk Σjm bm tk . (7.197)
One may write I
δWP K = fiP K d`δxi . (7.198)
C
In principle, the virtual work done by the self stress field can be also ex-
pressed by Eq. (7.196). However, in that case, Eq. (7.196) would become
a singular integral, which can be evaluated in the sense of Cauchy principal
value.
Since the core of a dislocation loop has specific physical meanings, it would
be appropriate to treat the virtual work of self-stress field separately. Gavazza
and Barnett [1976] expressed the virtual work of the self-stress field of planar
curved dislocation loop in terms of a single integral expression,
I h 8 i
00
δWself = E(t) − E(t) + E (t) ln κ − J(L, p) n · δxd`
C κ
+[dU ]core (7.199)
where E(t) = 12 σij (t)bi nj , is related to the core size, κ is the curvature of
the dislocation line, J(L, p) is a non-local interaction term, and [dU ]core is the
virtual work contribution from the core of the dislocation loop. Since [dU ]core
is related to the dislocation mobility, this term may be absorbed into the friction
force.
Let,
8 i
self 00
E = E(t) − E(t) + E (t) ln κ − J(L, p) (7.200)
κ
and
fiself = Eni (7.201)
The total active forces acting on a dislocation loop are
In many cases, it has to include the change of chemical potential induced Os-
motic force. Since the change in chemical potnetial per vacncy or interstitial
will cause the dislocation loop climbing, or causing the none-conservative dis-
location loop movement, the Osmotic force is usually responsible for the dis-
location loop climb (see Hirth, Rhee, and Zbib [1996]).
When a dislocation loop starting to move, it has to overcome the friction
forces that resist its motion. The friction forces consist of (1) extrinsic resis-
tances due to alloying, impurity atoms, Peierls stress (this part of force coming
162 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
from [dU ]core ), etc., and (2) Intrinsic friction forces that are due to the atom-
istic bond force in a surface separation (fracture) process. Empirically, one
can always assume that the friction forces are proportional to the dislocation
velocity, such that
I I
f riction
δW = Cik Vk d`δxi = C · V)d` · δx (7.203)
C C
where
dx
V= (7.204)
dt
and C is called the resistivity matrix, which has three independent components
in an isotropic medium (two for glide motion and one for climb motion),
C1 0 0
[Cik ] = 0 C2 0 (7.205)
C1 0 C3
Then the principle of virtual reads
I
int f ric
δW − δW = 0, ⇒ fiT − Cik Vk d`δxi = 0 . (7.206)
C
where Nim (u) is the finite element shape function. The discreteized velocity
field is
N
X DF
h h
Vi = xi,t = Nim (u)qm,t (t) . (7.209)
m=1
Denote the gradient of FEM shape function as Bim (u) := Nim,u (u). The
line integration element will be
NX
DF 1/2
d` = (x` x` )1/2 du = qp qs B`p (u)B`s (u) du (7.210)
p,s=1
Introduction of Dislocation Theory 163
We can evaluate the internal stresses acting on the dislocation loop by quadra-
ture integration, i.e.
Nloop Ns Qmax
I µ X X X h1
σij = bn wα R,mpp (jmn xi,u + imn xj,u )
4π 2
γ=1 β=1 α=1
1 i
+ kmn (R,ijm − δij R,ppm )xk,u (7.211)
1−ν
where Nloop is the total number of dislocation loops, Ns is the total number of
segments in each dislocation loop, and Qmax is the total number of quadrature
point in a segment, and wα is the quadrature weight.
Denote each segment of the dislocation loop as Lj . The discretized weak
formulation is
Ns QX
X max N
X DF h N
X DF i
Nim (u)δqm fiT − Cik Nkn q̇n
j=1 α=1 m=1 n=1
NX
DF 1/2
× qp qs B`p B`s wα = 0 . (7.212)
p,s=1
164 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Then, we can put the dislocation loop weak form into a matrix form,
Ns h
X h dq i iT h i
[f ]j − [γ]j δq = 0 , (7.215)
dt j j
j=1
dQ i T h i
h i h ih
F − Γ δQ = 0 , (7.216)
dt
where
h i h i1×NDF
F = AN s
j=1 f (7.217)
j
h i h iNDF ×NDF
Γ = AN s
j=1 γ (7.218)
j
where 0 ≤ α ≤ 1.
This is the state of the art discrete dislocation dynamics formulation.
One special case of great value for applications is that L real axis, Im(ψ(ζ)) =
g(x), Re(ψ(ζ)) = 0, Re(φ(ζ)) = f (x), and Im(φ(ζ)) = 0. That is φ(ζ) =
f (x) + i0 and ψ(ζ) = 0 + ig(x). Here f (x) and g(x) are real functions of
a real variable x satisfying the Holder condition for any finite x and vanishing
at infinity. This special case of Cauchy integral transforms is the so-called the
Hilbert transforms:
1 ∞ f (t)dt
Z
g(x) = H(f (x)) = (7.223)
π −∞ x − t
1 ∞ g(t)dt
Z
f (x) = −H(g(x)) = − (7.224)
π −∞ x − t
body displacement) ,
b h −1 y xy i
ux = tan + (7.228)
2π x 2(1 − ν)(x2 + y 2 )
b 1 − 2ν
h x2 i
uy = − ln(x2 + y 2 ) + (7.229)
2π 4(1 − ν) 2(1 − ν)(x2 + y 2 )
µb y(3x2 + y 2 )
σxx = − (7.230)
2π(1 − ν) (x2 + y 2 )2
µb y(x2 − y 2 )
σyy = (7.231)
2π(1 − ν) (x2 + y 2 )2
µb x(x2 − y 2 )
σxy = (7.232)
2π(1 − ν) (x2 + y 2 )2
As evident from the above equations, the stress fields are singular at the ori-
gin. Therefore the analytical solution presented above is no longer accurate
near the core of the dislocation. To remove this singularity inside the dislo-
cation core, Peierls [1940] and Nabarro [1947] included the discrete atomic
nature of the material and proposed the following lattice correction model.
The Peierls-Nabarro model(PN model) for a straight edge dislocation is de-
scribed using two semi-infinite simple cubic crystals as shown in Fig. 5.4. The
formal glide plane is y = 0. The two elastic half spaces are terminated on the
planes y ≥ d/2 and y ≤ −d/2. At the middle of glide plane, a non-Hookean
slab of width d (atomic spacing) joins the two half spaces. The symmetrical
configuration indicated in Fig. 5.4 suggests that this is done by cutting the
Introduction of Dislocation Theory 167
perfect crystal into two halves along the y = 0 plane, and inserting an addi-
tional layer of atoms in the upper half of the crystal space, which displaces the
upper half crystal moving rigidly a distance 0.5b in both positive and negative
x-direction, and we then re-weld the two half crystals.
Before the "re-welding", the initial dis-registry (misalignment) in x-direction
of two vertical atom layers with respect to the upper and lower half crystal
spaces is
b
2, x > 0
−
φ0x (x) := Xm+
− Xm = m = ±1, ±2, · · · ± ∞ (7.233)
b
− , x<0
2
After the re-welding, the misalignment, or the discontinuity, between the atom
layer in the upper part of crystal and the same atom layer (m) of the lower part
of the crystal becomes
φx (x) = x+ − x− + + − −
m = Xm + u (x) − (Xm + u (x))
m
b + −
2 + u (x) − u (x), x > 0
φx (x) =
b
− + u+ (x) − u− (x), x < 0
2
b
2ux (x) + 2 , x > 0
=
b
2ux (x) − , x < 0
2
By antisymmetry, we assume that ux (x) = u+ (x) = −u− (x).
At the remote boundary, dis-registry is enforced to be zero, i.e. there is no
discontinuity at the remote boundary
b
φx (x) → 0, when x → ±∞ ⇒ 2ux (x) ± = 0, x → ±∞ (7.234)
2
b
Therefore, ux (±∞) = ∓ . This implies that the total displacement along the
4
interface should be
Z ∞
dux b
ux (∞) − ux (−∞) = dx0 = − (7.235)
−∞ dx x=x 0 2
Based on Eshelby’s interpretation (Eshelby [1949]), one may think that
Peierls-Nabarro model deploys a continuous edge dislocation distribution along
168 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
the cohesive interface with its local Burgers vector density as b0 (x0 ) to replace
a single dislocation with a Burgers vector b. To make sure that these two
dislocation systems are equivalent, we enforce the following condition on net
Burgers vector equality,
Z ∞ Z ∞
dux 0
−2 dx = b0 (x0 )dx = b (7.236)
−∞ dx x=x 0
−∞
From the above relation, one may derive that the distribution density of Burgers
dux 0
vector should be b0 (x0 ) = −2 (x ).
dx
The strains near the dislocation core are large, and therefore use of Hooke’s
law for the stresses is unappropriate. One the other hand, it is relevant to use
the periodicity of the lattice, which implies σxy to be a periodic function of
φ(x). We therefore assume that,
2πφ
x
σxy (x, 0) = C sin (7.237)
b
2πφx (x)
When φx (x) << 1,σxy (x, 0) ∼ C . Under small deformation limit,
b
it is assumed that the cohesive law should comply to Hooke’s law as well (is
this a good assumption?), i.e.
φx (x)
γxy = (7.239)
d
Thereby, one obtain that
One may also derive the above integral equation based Boussinesq solution of
linear elastic half space (e.g. Timoshenko and Goodier [1972]).
Apparently, σxy (x, 0) is proportional to the Hilbert transform of dux /dx.
Thereby the inverse Hilbert transform gives
(1 − ν) ∞ σxy (t, 0)dt
Z
dux
= (7.242)
dx µ −∞ x−t
Integrating this yields,
∞
(1 − ν)
Z
u(x) = σxy (t, 0) ln |t − x|dt (7.243)
µ −∞
Using ((7.240)) and ((7.241)), one can obtain the well-known Peierls-Nabarro
integral equation for unknown displacement field, ux (x),
Z ∞
(dux /dx)x=t dt b(1 − ν) 4πux
= sin (7.244)
−∞ x−t 2d b
which is a singular, nonlinear integral equation with unknown function ux (x).
Luckily, the solution of the above integral equation can be found in closed
form 3 ,
b x
ux (x) = − tan−1 (7.245)
2π rc
where rc = d/2(1 − ν), which is a parameter that characterizes the size of the
dislocation core. When |x| < rc , the dis-registry φx (x) > b/4. At x = rc ,
ux (rc ) = −b/8 and φx (rc ) = b/4.
Substituting ((7.245)) into ((7.240)) and utilizing the trigonometry identity
y
tan−1 (y) = sin−1 p
1 + y2
one can find that
µb x
σxy (x, 0) = (7.246)
2π(1 − ν) x + rc2
2
3 My guess is that the reason why they took sine function as the cohesive law was to match the exact solution
µb 3 4πux
= 2
1 + cos (7.252)
8π d b
Introduction of Dislocation Theory 171
Substitute,
b x
ux = − tan−1 (7.253)
2π rc
to obtain the misfit energy for a pair of atomic planes as,
µb3 x
∆W = 2 1 + cos 2 tan−1 ( ) (7.254)
8π d rc
Let the distance of the center of the dislocation from the nearest position of
symmetry be ξ = αb, where α is a variable. Then the position of all the atoms,
on the two faces of the slip plane are defined by
2m b
the upper half crystal
xm = 2 (7.255)
b
(2m − 1)
the lower half crystal
2
and m = 0, ±1, ±2, ±3, · · · (see Fig. 7.14).
Then the total misfit energy is the summation,
∞
X
W = ∆W (2m) + ∆W (2m − 1)
m=−∞
··· +∞
X µb3 X
−1 b
= 1 + cos 2 tan (α + 0.5n)( )
8π 2 d n=−∞ rc
n=0,±2,±4
··· +∞
X µb3 X
−1 b
+ 1 + cos 2 tan (α + 0.5n)( ) (7.256)
8π 2 d n=−∞ rc
n=±1,±3
where we have used the fact that the function f (n) is even in n. We can rewrite
the above relation as,
+∞
X Z +∞ ∞ Z
X +∞
f (n) = f (x)dx + 2 f (x)cos(2πxn)dx, (7.260)
n=−∞ −∞ n=1 −∞
Therefore we can rewrite the total misfit energy from the equation ((7.257))
as,
+∞
µb3
Z
W = (1 + cos(2 tan−1 z))dx
8π 2 d −∞
+∞ Z
µb X +∞
3 dz
+ (1 + cos(2 tan−1 z))cos 2πn − 2α dx
4π 2 d (1 − ν)b
n=1 −∞
(7.261)
+∞ Z +∞
µb2 X dz dz
+ cos 2πn − 2α
2π 2 (1 − ν) (1 − ν)b 1 + z2
n=1 −∞
(7.262)
The first integral above can be calculated using the Cauchy residual theorem,
that is we use the result:
Z +∞
1 1
2
dz = 2πiRe( )=π
−∞ 1 + z 1 + z2
where Re(.) denotes the residual. Therefore the first term of the total misfit en-
µb2
ergy as 4π(1−ν) . The second term in equation ((7.262)) can be further reduced
to,
+∞ Z +∞
µb2 X 2πnzd dz
cos(4πnα) cos
2π 2 (1 − ν) −∞ (1 − ν)b 1 + z 2
n=1
2πnd
To evaluate this term we again use Cauchy residual theorem. Say k = (1−ν)b ;
then the integral in the above equation is equal to,
Z +∞ ikz
e
2
dz
−∞ 1 + z
which is equal to πe−k . Therefore we obtain the total misfit energy as,
+∞
µb2 µb2 X −4πrc n
W = + 2 πe b cos(4πnα) (7.263)
4π(1 − ν) 2π (1 − ν)
n=1
1 dW (α)
F =− (7.265)
b dα
Note that the dislocation moves a distance −αb.
174 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
dW (α)/dα reaches to maximum when sin 4πα = 1. From the relation that
σxy = F (b × 1)(unit thickness in z-direction), the critical shear stress to move
the dislocation by one lattice site is
2µ 4πr
c
σ= exp − (7.266)
(1 − ν) b
where F is called the Peierls force and σ is called the Peierls stress, which are
required to move a dislocation over a Peierls barrier.
A more physically realistic restoring stress is obtained if we use relative
displacement (of the two half planes) instead of the lattice displacement in the
above discussion. In the following, a more recent treatment of the PN model is
outlined (Joós and Duesbery, 1997) which considers the relative displacement
instead of the independent lattice displacements in two half planes. We restrict
our attention to the case of a straight edge dislocation. The new model predicts
a Peierls stress which differs from the above mentioned expression by a factor
of two in both the exponential and the coefficient of the exponential. This
approach is also valid for the case of narrow dislocations. By f (x) we define
the displacement of the upper half of the crystal with respect to the lower half.
If c is a constant, then f (x − c) corresponds to a dislocation translated by
c. For a discrete lattice this can be understood like this: If the dislocation is
introduced at c, then the atomic planes at a position mb in the upper half of the
crystal will experience a displacement of f (mb − c) along the Burgers vector.
The total misfit energy in this case can be written as:
+∞
µb3 X
−1 mb − c
W (c) = 1 + cos 2 tan ( ) (7.267)
4π 2 d m=−∞ rc
Note the difference of factor of half in the expression of W from the earlier dis-
cussion. This is because we are no longer treating the two half planes indepen-
dently, but we are using a relative displacement. Using further manipulations
and substituting Γ = rc /b and y = c/b we have,
+∞
µb2 X Γ
W (y) = (7.268)
4π (1 − ν) m=−∞ Γ + (m − y)2
2 2
Where we can calculate the Fourier coefficients in the usual manner. After
substituting the value of these Fourier coefficients, we get the expression for
Introduction of Dislocation Theory 175
For the limit of wide dislocations (Γ 1), only the first exponential term is
kept. Then in the limit of wide dislocations we have,
µ 2πr
c
σ= exp − (7.272)
(1 − ν) b
Note the difference between the above stress and the one obtained in the equa-
tion ((7.266)).
Figure 7.15. Paul Dirac (left), Wolfgang Pauli (middle) and Rudolf Peierls (right) in discus-
sion at the international Conference on Nuclear Physics, Birmingham, 1948
"1937 I was invited to work at the University of Birmingham, in the Physics Department
which had just taken over by M. L. E. Oliphant (now Sir Mark Oliphant). I felt that it
would be urgent to know the width of the dislocation belt and the stress required to
move it. The simplest assumption about this was the one made by Taylor, that the
stress was zero; however, the extremely high yield stress of many hard materials such
as diamond (which could be remarkably free from imperfections and thus could not
contain too many dislocations) indicated that the most frequent cause of the hardness
of crystalline materials was the high shear stress required to move a dislocation. I found
that the width of the dislocation and the stress for moving it could be calculated, with
a crude approximation, simply enough by assuming that the shearing force between
the opposite shores of the slip plane in a dislocation was a sine function of the relative
shear displacement (the initial tangent of the sine, of course, was given by the elastic
modulus).
One the other hand, displacement and shear traction at the surface of a half-space were
connected by the equations of Boussinesq; equating the stresses and displacements of
the sine approximation with those of Boussinesq led to an integral equation which was
the solution of the problem It would have taken me days or weeks of study to solve
it; fortunately I was a daily guest in the hospitable house of the brilliamnt theoretical
physicist Rudolf Peierls. He solved the equation, if I remember well, within a few
hours, and he also drove me to a conference at Bristol University in 1939 where I gave
a paper and he gave another on the problem he had just solved.
The calculation of the width of the dislocaiton and of the Peierls-Nabarro stress required
for moving it was repeated and improved by Nabarro in 1947. The result was puzzling
at first: the width calculatied by Nabarro amounted to a few atomic spacings while
Peierls obtains an order of magnitude of thousands of spacings. After some research in
Birmingham and in Cambridge (where I was wat the time) I discovered the sheet with
Peierls’s calculations in my desk; Peierls checked it and found that a factor of 2π was
accidentally omitted in an exponent, which amounted to a factor of about 1000 in the
result.
Of course, the calculation with the sinusoidal approximation is useless in most interest-
ing cases of directinal bonds, in transition metals and the hard non-metallic crystals."
From The Sorby Centennial Symposium on the History of Metallugy, MSC,
Vol. 27, 1963, pages 368-369.
defect-free epitaxial thin films has been the main challenge in semi-conductor
industry in the past half century.
In this section, we shall introduce the two basic dislocation models in thin-
film mechanics.
7.8.1 Frenkel & Kontorova model and Frank & van der
Merwe model
The Frenkel & Kontorova dislocation model is a one-dimensional disloca-
tion model, which was proposed in 1937. This model was studied in detailed
by Frank and van der Merwe [1950ab], and they applied it to study thin film
mechanics or epitaxial thin film mechanics.
In Frenkel & Kontorova model, the thin film is modeled as one dimensional
monolayer with lattice spacing af , and the substrate is modeled as large slab
with lattice spacing as , and as 6= af and the lattice misfit is ∆ = af − as (see
Fig. 7.17).
The row of atoms in the thin film are under combined influence of harmonic
forces between the nearest neighbours in the monolayer and non-linear inter-
action forces from substrate. Since the substrate is assumed much larger in
dimension than the thin film, it is assumed to be rigid. The interaction between
the thin film and substrate, or the force exerted on the thin film by the sub-
178 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
After attach the thin film onto the substrate, the thin film will be stretched to
the position
where xrm is denoted as the reference position of the m-th atom with respect to
the aubstrate, and umis
m is the displacement of the atom due to the lattice misfit,
umis
m = m(as − af ) .
xm = Xm + umis e
m + um (7.275)
or
um = xm − xm = umis e
m + um (7.276)
where uem is the elastic deformation of the atom.
The relative displacement between the two atoms is now
2πuem
1X e e 2
Π= µ(um+1 − um − (af − as )) + W [1 − cos ] (7.278)
2 m a
Let,
uem af − as
ξm = , and f = . (7.279)
as as
Hence
1 X 2
Π= µa (ξm+1 − ξm − f )2 + W [1 − cos(2πζm )] (7.280)
2 m
i.e.
π
∆2n ξ = (ξn+1 − 2ξn + ξn−1 ) = sin 2πξn (7.282)
2`20
p
where `0 = µa2 /2W .
The dynamics version of Eq. (7.282) is the finite-difference sine-Gordon
equation,
mn d2 ξn π
∆2n ξ − = 2 sin 2πξn (7.283)
µ dt2 2`0
If `0 >> 1, one may use continuous approximation to replace the finite
difference equation with a differential equation,
d2 ξn 2 2 d4 ξ 4 d2 ξ
∆2n ξ = a + a + O(a6
) = + O(a4f ) (7.284)
dXn2 f 4! dXn4 f f
dn2
Therefore, if we only consider static deformation, we have the following non-
linear ordinary differential equation
d2 ξ π
2
= 2 sin 2πξ . (7.285)
dn 2`0
= 0, and k = 1. (7.291)
π n
Z Z ξ
dζ
dp = π (7.293)
`0 0 0 sin πζ
where the upper limit φ is called the amplitude. The inverse relation of the
above elliptic function is πn
φ = am (7.298)
`0 k
or
1 1 πn
ξ = + am (7.299)
2 π `0 k
and
dξ 1 πn 1
= dn = (1 − k 2 cos2 πξ)1/2 (7.300)
dn `0 k `0 k `0 k
At ξ = ξ(0) = 1/2,
dξ 1
= (7.301)
dn `0 k
i.e. `0 k is now the effective dislocation length.
182 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
dξ
Figure 7.19. The general solution of static sine-Gordon equation ( dn ≥ 0).
Assume that ξ(p) = 1.5. The general solution of FKV model is depicted on
Fig. 7.19. Obviously, the number is the atoms per dislocation,
2`0 kE(k)
p= (7.302)
π
where E(k) is the following elliptic integral,
Z π/2
E(k) = (1 − k 2 sin2 ψ)1/2 dψ (7.303)
0
The general solution indicates that there are many dislocation occuring simu-
tanelously along the chain in periodic fashion. In Fig. 7.20, we show the
dislocation pattern created by the general solution.
It would be interesting to examin the stability of Frenkel-Kontorova system.
The potential energy of one dislocation
p−1 p−1
X W X
Π = W `20
(ξn+1 − ξn − f ) + 2
1 − cos 2πξn )
2
n=0 n=0
Z P Z P
2 dξ 2
= W `0 − f dn + W sin2 πξdn (7.304)
0 dn 0
Consider
dξ 1
= 2 2 (1 − k 2 cos2 πξ)1/2 (7.305)
dn `0 k
Introduction of Dislocation Theory 183
dξ
Figure 7.21. Dislocation pattern for dn
≤ 0.
When k = 1,
1 2 W 1/2
fcr = = (7.309)
p π µa2 /2
It is beleived that when lattice misfit f > fcr , dislocations will spontaneous-
lly enter or depart from the monolayer chain.
We can find a critical thickness, hcr , of the thin film below which the thin film
will stay in a coherent state with the substrate that is the thin film is defect-free.
From (7.315), one can find that the critical thickness can be determined from
the following non-linear equation,
hcr µb
βh = (7.316)
cr 4π(1 − ν)M
ln
b
Exercise
Probelm 7.1 Consider cuboidal region of inelastic strain (eigenstrain) due
to solute segregation forming cuboidal precipitates. The precipitate subdomain
(or inclusion) has the dimension 2a × 2a × 2a, and the unit cell (U) has the
dimension 2L × 2 : ×2L. The eigenstrain is assumed to have a constant value
within each inclusion, and be zero outside the inclusion,
δij ε, ∀ x ∈ Ω;
ε∗ij = (7.317)
0; ∀ x ∈ U/Ω
where
n o
U =x − L ≤ xi ≤ L, i = 1, 2, 3 (7.318)
n o
Ω = x − a ≤ xi ≤ a, i = 1, 2, 3 , and a < L (7.319)
Chapter 8
Assume that p(x), q(x), and f (x) are given continuous functions, i.e. p(x), q(x),
and f (x) ∈ C 0 [x0 , x1 ], and p(x) > 0, q(x) > 0. Let,
We usually call y as the trial function and αη(x) as the test function.
Comparison Variational Principles 189
In order to find the function y(x) that yields the extreme value of I[y], we
consider the value of I[ỹ],
Z x1 n
0 0
I[y(x) + αη(x)] = p(x)[y (x) + αη (x)]2 + q(x)[y(x) + αη(x)]2
x0
+2f (x)[y(x) + αη(x)]} dx
Z x1 h i
0
= p(x)(y (x))2 + q(x)y 2 (x) + 2f (x)y(x) dx
x0
Z x1 h i
0 0
+ 2α p(x)y (x)η (x) + q(x)y(x)η(x) + f (x)η(x) dx
x0
Z x1
2 0
+α p(x)(η (x))2 + q(x)η 2 (x) dx (8.7)
x0
Thereby,
α2 2
∆I = I[y(x) + αη(x)] − I[y(x)] = αδI + δ I (8.8)
2!
where
Z x1
0 0
δI = 2 [p(x)y (x)η (x) + q(x)y(x)η(x) + f (x)η(x)]dx (8.9)
Zx0x1 h i
0
δ2I = 2 p(x)(η (x))2 + q(x)η 2 (x) dx (8.10)
x0
We say that
I[y] is stationary at y = y(x) if δI = 0. Since both p(x), q(x) > 0
y=y(x)
and δ 2 I > 0, I[y] will reach a minimum at y = y(x).
The first order variation illustrated above is in the sense of Gateaux. The
definition of the Gateaux variation is in terms of the so-called Gateaux deriva-
tive
I(y + αη) − I(y) d
δG I = DG Iη = lim = I(y + αη) (8.11)
α→0 α dα α=0
Remark 8.1.1 One may compare this with the so-called Fr«echet derivative,
DF I[y]η, which is defined as a linear functional such that
I(y + η) − I(y) − DF I(y) · η
⇒ 0, as kηkV → 0 . (8.12)
kηkV
Gateaus derivative coincides with Fre’chet derivative, if δF I is linear in η and
uniformly continuous in η, i.e. |δI(y, η)−δI(y0 , η)| → 0, as y → y0 uniformly
∀y ∈ B(y0 ).
190 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
n dn
δG I= I(y + αη) , ∀n ≥ 1 (8.13)
dαn α=0
such that
α2 2 α3 3 α4 4
∆I = I(y + αη) − I(y) = αδG I + δG I + δG I + δG I + · · · (8.14)
2! 3! 4!
In the rest of the book, we omit the subscript G in variation operator. Let
α = 1. We have
1 2 1 1
∆I = I(y + η) − I(y) = δI + δ I + δ3I + δ4I + · · · (8.15)
2! 3! 4!
One nice thing about the Gateaux variation is that it is defined based on a
scaler differentiation operation. In other words, the variation operation follows
the same rule as the differentiation operation in elementary calculus.
This can be seen by examining the first order variation of I[y],
Z x1 h i
0 0
δI = 2 p(x)y (x)η (x) + q(x)y(x)η(x) + f (x)η(x) dx (8.16)
x0
This is to say that one can find the first variation of a functional, I[y], by
simply differentiating (taking G-derivative) the unknown function according
to the same rule of differentiation in calculus. The only difference is: dy is
replaced by δy, which is the variation of the unknown function, or in general,
◦
a test function satisfying homogeneous boundary conditions, i.e. δy ∈ V.
Consider the first term in (8.16). Integration by parts yields,
Z x1 Z x1 Z x1
0 0 0 x1 0 0 0
p(x)y η dx = [p(x)y η]x0 − (p(x)y )ηdx = − (p(x)y ) ηdx
x0 x0 x0
Therefore,
Z x1 h i
0 0
δI = 2 −[p(x)y (x)] + q(x)y(x) + f (x) η(x)dx = 0 (8.17)
x0
Comparison Variational Principles 191
◦
Since this equation must holds for any η(x) ∈ V, the integrand must vanish,
i.e. the solution of the following differential equation
0 0
−[p(x)y (x)] +q(x)y(x)+f (x) = 0, y(x0 ) = y0 and y(x1 ) = y1 . (8.18)
and for this purpose we call a function that makes I[y] stationary, but not nec-
essarily satisfy the Euler-Lagrange equation, i.e.,
Z x1 h i
0 0
δI = 2 p(x)y (x)δy + q(x)y(x)δy + f (x)δy dx (8.20)
x0
∂F ∂ ∂F
E[F ]y = − =0. (8.23)
∂y ∂x ∂y 0
192 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
◦
and test function δui ∈ V where,
◦ n o
V := ηi (x) ηi (x) ∈ H 1 (V ), and ηi (x) = 0, ∀x ∈ Γu (8.27)
by parts,
Z Z Z
δΠ = σij δui,j dV − fi δui dV − t0i δui dS
ZV V
ZΓt Z
= (σij δui ),j − σij,j δui dV − fi δui dV − t0i δui
V V Γt
Z Z Z
= σij nj δui dS − (σij,j + fi )δui dV − t0i δui dS
Z∂V V
Z Γt
Z
0
= (σij nj − ti )δui dS − (σij,j + fi )δui dV + σij nj δui dS
Γt V Γu
u = x · 0 , x ∈ ∂V (8.33)
which is a map,
Πc : S → IR (8.38)
where S is the trial function space
n o
S = σij σij ∈ H 1 (V ), σij,j = 0 and nj σij = t0i , ∀x ∈ Γt (8.39)
Note that in this variational statement, the essential boundary condition be-
comes
nj σij = t0i , ∀x ∈ Γt (8.41)
whereas the natural boundary condition becomes
ui = ūi , ∀x ∈ Γu . (8.42)
The necessary condition for Πc (σij ) attaining extreme value is the stat-
tionary condition,
δΠc = 0 .
196 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Hence
1 2 c
∆Πc = δ Π >0 (8.43)
2!
since Dijk` is positive definite. Thus, Πc (σij ) reaches a minimum value at
σij = σ̃ij , where σ̃ij renders stationary condition δΠc (σ̃ij ) = 0. This fact is
the so-called minimum complementray potential energy principle.
Theorem 8.2 (Minimum Complementary Energy Principle) Among
all statically admissible stress fields, the actual stress field (whose correpond-
ing strain field satisfies compatibility condition) rensers Πc an absolute mini-
mum, i.e.
Πc (σ̃) ≤ Πc (σ), ∀σ ∈ S (8.44)
or
Πc (σ̃) = inf Πc (σ) (8.45)
σ ∈S
The stationary condition of complementary energy has well-known names,
e.g. virtual force principle in continuum mechanics, or the weak form of com-
patibility condition in computational mechanics,
Z Z
c
δΠ (σ̃ij ) = Dijk` σ̃ij δσk` dV − u0i δσij nj dS = 0 (8.46)
V Γu
ui = u0i , ∀x ∈ Γu (8.49)
< σ >= σ 0 , ∀ σ ∈ S .
Choose σ = σ 0 ∈ S,
Z Z
1 1
Wc (σ) = σ : dV = σ 0 : dV
2V V 2V V
n
Z X
1
= σ0 : Dα : σ α dV
2V V α=0
n
1 0 X Ωα α
= σ : D : σ0
2 V
α=0
n
1 0 X
= σ : fα Dα : σ 0
2
α=0
198 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Therefore,
n
X
σ 0 : D̄ : σ 0 < σ 0 : fα Dα : σ 0 (8.54)
α=0
Since D̄ : C̄ = 1(4s) and both D̄ and C̄ are positive definite, we then have
n
X −1
fα C−1
α ≤ C̄ (8.55)
α=0
which is called the Reuss bound. It is a lower bound for elastic moduli.
Assume that
and consequently,
n
1 X
n < K̄ < fα Kα
X fα
α=0
Kα
α=0
n
1 X
n < µ̄ < fα µα
X fα
α=0
µα
α=0
One can see that the Voigt bound is in fact an arithmetic average and the Reuss
bound can be viewed as a geometric average or the harmonic average.
Comparison Variational Principles 199
σij,j = 0,
σij = Cijk` (x)k` ,
1
U () = Cijk` ij k` , and W () =< U () >V
2
ui = ūi , ∀x ∈ Γu , (Γt = ∅, Γu = ∂V ).
and
(0)
σij = pij + Cijk` k`
(0) (0)
= pij + Cijk` (ijk` + dijk` ) (8.58)
where udi is the disturbance displacement field and pij is called polarization
stress.
A better definition of stress polarization is
(0) (0)
pij = σij − Cijk` k` = (Cijk` − Cijk` )k` (8.59)
200 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Π : S × U → IR,
where
Z
1 (0) (0) (0) −1 (0)
Π(pij , dij )
= d
Cijk` ij k` − ∆Cijk` pij pk` + pij ij + 2pij ij dV
2
V
(0)
∆Cijk` = Cijk` − Cijk`
where pij = ∆Cijk` k` (8.67)
d
(0)
ij = ij − ij
2.
Proof:
which leads to
(0)
δtij = δpij + Cijk` δdk` , and δtij,j = 0 . (8.73)
Substituting (8.72) and (8.73) into (8.71) yields
1 Z
(0) (0)
δΠ = − dij (δtij − Cijk` δdk` ) − δdij (tij − Cijk` dk` ) dV
2 V
1 Z
d
(0) (0)
ij δtij − tij δdij − dij Cijk` δdk` + δdij Cijk` dk`
= − dV
2 V | {z }
=0, because C(0) has major symmetry
1 Z
= − udi,j δtij − tij δudi,j dV
2 V
Considering the facts
Z Z Z
δtij udi,j dV = δtij nj udi dS − δtij,j udi dV ≡ 0
V
Z Z∂V ZV
tij δudi,j dV = tij nj δudi dS − tij,j δudi dV ≡ 0,
V ∂V V
(0)
Substitute δpij = δtij − Cijk` δdk` into (8.75). It can be readily shown that
Z
(0) −1
(0)
I = Cijk` δtij δtk` − 2δtij δdk` +Cijk` δdij δdk` dV
V | {z }
=0
Z
(0) −1
(0)
= Cijk` δtij δtk` + Cijk` δdij δdk` dV
V
A direct consequency is
Z Z
(0) −1 (0)
Cijk` δpij δpk` dV > Cijk` δdij dk` dV (8.76)
V V
It is clear now that if ∆C−1 < 0, δ 2 Π > 0 and hence Π has a global minimum.
To sum up, we have the following extreme conditions,
♣
204 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
because udi
= 0, ∀x ∈ ∂V .
Therefore the total potential energy of a kinematically admissible field, ui ∈
V, can be written as
Z Z
1 1 (0)
Π() = σij ij dV = σij ij − σij dij dV
2 V 2 V | {z }
=0
Z
1 (0)
= σij ij dV
2 V
Consider
(0) (0) (0) (0)
σij ij = σij + pij + Cijk` dk` ij
(0 (0) (0) (0) (0) (0) (0)
= σij ij + pij ij + Cijk` dk` ij + + pij ij − pij ij
| {z }
=0
(0) (0) (0) d (0) (0)
= Cijk` k` ij + Cijk` k` ij + +2pij ij − pij (ij − dij )
(0) (0) (0) (0) (0) (0)
= Cijk` k` ij + Cijk` ij dk` +2pij ij − pij ij + pij dij
| {z }
=0
u + (v + w) = (u + v) + w, ∀u, v, w ∈ V
2 Commutivity of addition
u + v = v + u, ∀u, v ∈ V
0 + u = u + 0 = u, ∀u ∈ V
u + (−u) = (−u) + u = 0
r(u + v) = ru + rv
(r + s)u = ru + rv
rsu = r(su)
1u = u
Remark 8.4.1 1 The first four properties in the definitions of vector space
can be summarized that V is an abelian group under addition;
2 Any expression of the form
r1 v1 + r2 v2 + · · · + rn vn
r1 v1 + r2 v2 + · · · + rn vn ∈ V
206 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
V × V → V : (u, v) → u + v ∈ V
F × V → V : (α, u) → αu ∈ V
are closed.
5 When the operations
f : (u, v) → u + v ∈ V
g : (α, u) → αu ∈ V
are continuous, the vector space is called topological vector space.
Example 8.5 Let F = IR. The set of all ordered n-tuples, i.e.
u = (u1 , u2 , · · · , un ), ui ∈ IR
and
α(a1 , · · · , an ) = (αa1 , · · · , αan )
is a vector space, and it is denoted as IRn . Note that in general vector space
(a mathematical concept) is still a primitive set. It may have some algebraic
structures, but it does not have topologival structures, or geometric structures,
such as distance between two elements.
Example 8.6 Let F = IR. The set of all continuous function, C 0 (IR), i.e.
∀f ∈ C 0 (IR)
f : X ⊂ IR → Y ⊂ IR
and
dY (f (x), f (y)) < , ∀dX (x, y) < δ, ∀δ > 0 .
is a vector space under the operations of addition and scalar multiplication,
i.e.
(f + g)(x) = f (x) + g(x), f, g ∈ C 0 (IR)
and
αf (x) = αf (x), ∀α ∈ IR, f ∈ C 0 (IR)
Comparison Variational Principles 207
(·, ·) : X × X → IR
with properties:
1 (x, x) ≥ 0, ∀x ∈ X and (x, x) = 0 iff x = 0;
3 Linearity
(αx + βy, z) = α(x, z) + β(y, z),
and
This particular inner product space is denoted as En = {IRn , (·, ·)}. It gener-
ates a norm,
Xn 1/2 p
kxk`2 := xi xi = (x, x)
i=1
208 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
This norm is called Euclidean norm on IRn . The space is therefore a normed
space as well — called n-dimensional Euclidean space, En = {IRn , k · k`2 }.
One can show that
(i) kxk`2 ≥ 0, ∀x ∈ En
kxk`2 = 0, ⇐⇒ x = 0 ;
(ii) kαxk`2 = |α|kxk`2 , ∀xEn , α ∈ IR
(iii) kx + yk`2 ≤ kxk`2 + kyk`2 ← triangle inequality;
(iii) k(x, y)k`2 ≤ kxk`2 kyk`2 ← Cauchy − Schwartz inequality;
Based on the `2 -norm, one can measure the distance between two vectors in
En ,
ρ(x, y) := kx − yk`2 ;
One can also show that
(i) ρ(x, y) = ρ(y, x);
(ii) ρ(x, y) > 0, and ρ(x, y) = 0, iff x = y;
(iii) ρ(x, y) ≤ ρ(x, z) + ρ(z, y), ∀x, y, z ∈ En
The distance function ρ(x, y) is called a metric, and the associated vector
space is called metric space.
Example 8.10 (L2 Space) Consider a real value function f (x), x ∈ [a, b].
Define an inner product,
Z b
(f, g) = f (x)g(x)dx
a
Therefore, L2 ([a, b]) is an inner product vector space, and of course, normed
space (metric space).
Example 8.11 (Lebesgue Space (Lp (Ω))) Let Ω be an open set in IRn .
For 1 < p < ∞, one can define a Lp -norm for a measurable function f ,
Z 1/p
kf kLp (Ω) := |f (x)|p dx
Ω
k
X 1/p
kf kWpk (Ω) = kDα f kpLp (Ω)
α=0
Note that the Sobolev norm is not generated by an inner product in general.
A Sobolev space is defined as
and sZ
h i
kf kH 1 (Ω = f (x)2 + ∇f (x) · ∇f (x) dV
Ω
and
sZ
h i
kf kH 2 (Ω = f (x)2 + ∇f (x) · ∇f (x) + ∇ ⊗ ∇f (x) : ∇ ⊗ ∇f (x) dV
Ω
Comparison Variational Principles 211
Define
Z
U : E → IR, U () =
σ(˜
)d˜
0
Z σ
c ∗ c
U : E = S → IR, U (σ) = (σ̃)dσ̃
0
Both strain energy density and complementary strain energy density are con-
vex, and they are plotted in Fig. (8.4).
Comparison Variational Principles 213
8.4.2 G^
ateaux variation and convex functional
The G^ateaus variation of a functional in a linear space is the generalized
directional derivative of a real-value function in vector calculus.
Note that
d
δP (ū, u) = P (ū + λu)
dλ λ=0
δP
:= DP (ū)
δu
Question: why are convex functionals so special ? The following theorem
answers this question:
Theorem 8.19 If P : Uk ⊂ U → IR is G^ateaux differentiable, then, the
following statements are equivalent to each other
(S1) P : Uk ⊂ U → IR is convex;
(S2) P (v) − P (u) ≥< v − u, DP (u) >, ∀v, u ∈ Uk
(S3) < v − u, DP (v) − DP (u) >≥ 0 , ∀v, u ∈ Uk
Remark 8.4.3 The statement (S3) shows that G^ateaux derivative of a con-
vex function is a monotone operator of U into U ∗ . By the mean value theorem,
D2 P (u) ≥ 0, ∀u ∈ Uk
This is to say that if elastic tensor is positive definite, the elastic potential
energy is convex. Similar statement can be made for complementary potential
energy, if the compliance tensor is positive definite.
3 The stationar (or sta) primal variational problem is to find a stationary point
u ∈ Uκ such that
Psta : P (ũ) = sta P (u), ∀u ∈ Uκ
Remark 8.4.4 1 A stationary point is also called critical point. The criti-
cal point condition,
δP (ũ, u) = 0, ∀u ∈ Uκ
Proof:
Comparison Variational Principles 217
We only prove the theroem in IR2 , which has the full flavor of a rigorous
proof.
We first show (T1). The tangential vector from the pole to the paraboloid is
t = (σ̄ − , Y (σ̄) − X)
1
the normal vector of graph G = U − 2 = 0 is
2
∂G ∂G
n= , = (−, 1)
∂ ∂U
We want show that the contact point is in the polar : X() = σ̄ − Y (σ̄).
Consider the condition t · n = 0.
t · n = (σ̄ − , Y − X)(−, 1)
= −σ̄ + 2 + Y − X
= −σ̄ + 2X + Y − X = −σ̄ + X + Y = 0
(D1) A regular point of the function U () is a point ∈ E where the deter-
∂2U
minant of the Hessian matrix D2 U = { } satisfies,
∂i ∂j
2
∂ U
det 6= 0, or ± ∞
∂i ∂j
(D2) A regular domain, denoted by Er is a continuous subset of regular
points.
∂2U ∂2U c
= δij .
∂i ∂k ∂σk ∂σj
The proof of this theorem is basically application of implicit function theo-
rem. It is omitted here. The readers who are interested in the proof may consult
Gao [2000].
Now we move to the essentail technical ingradient of convex analysis.
Theorem 8.24 (Duality between the regular manifolds) Let U
and U c be Legendre dual functions over the duality domain E and E ∗ respec-
tively.
Comparison Variational Principles 219
U c = min{σ − U ()}
∈E
Proof;
For simplicity, we only prove it for case E ⊂ IR, which contains the enssen-
tial substance of a general, rigorous proof.
∂U
Since σ = , by Taylor expansion,
∂
∂U ∂2U
σ= + 2 ∗ ( − ¯) (8.86)
∂ =¯ ∂
where
∂2U ∂2U
∗ =
∂2 ∂2 =¯
+θ∆
and 0 ≤ θ ≤ 1.
220 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
∂2U
(σ − σ̄) = + ∗ ( − ¯) (8.87)
∂2
∂U c
By the same token, because of = , one can have
∂σ
∂2U
( − ¯) = + 2 ∗ (σ − σ̄) (8.88)
∂
where
∂2U c ∂2U
∗ =
∂σ 2 ∂σ 2 =σ̄+θ∆σ
and 0 ≤ θ ≤ 1. Therefore,
∂2U
(σ − σ̄)( − ¯) = ∗ ( − ¯)2
∂2
∂2U c
= ∗ (σ − σ̄)2 (8.89)
∂σ 2
∂2U ∂2U c
Eq. (8.89) indicates that if 2
∗ is positive definite, 2
∗ is also
∂ ∂σ
∂2U ∂2U c
positive; whereas if ∗ is negative definite, ∗ is also negative
∂2 ∂σ 2
definite, or both being indefinite.
To prove the Legendre inequality, we consider a special 1D example, U () =
1 2
k0 , k0 > 0.
2
For a given point ¯ on horizontal axis, the associated stress σ̄ = k0 ¯ is the
slope of the polar, the straight line X = σ̄ − Y , which is tangent to the graphy
of U at ¯ (see Fig. (8.7).
Therefore, point (¯ ) is in both polar X = σ̄ − Y and on U = 1/2k0 2 ,
, U (¯
which is to say that X(¯ ) = U (¯ ) and
) =: U c (σ̄)
− U (¯
Y = σ̄¯
∂2y
Since U () is convex, y() is then concave because < 0. It then takes its
0
∂2
maximum value at ¯ because y (¯
) = 0. That is
is used.
Therefore,
Z Z Z
1 1
Cijk` ˜ij ˜k` dV ≥ σij 0ij dV − Dijk` σij σk` dV
2 V V 2 V
which is essentially
where
Z
1
W () = Cijk` ij k` dV
2V V
Z
1
W c (σ) = Cijk` σijσk` dV
2V V
One may further tighten the bound
0
W (˜
) = sup :< σ > −W c (σ̃) (8.92)
{<σ >:σ ∈S}
Remark 8.6.1 1. Note that Eq. (8.91) looks like Legendre-Fenchel trans-
formation. However, there is a subtle difference.
If W is a convex functional of ∈ E, the Legendre-Fenchel transformation
assures that
W c (σ) = sup {σ : − W ()}
{∈E}
2. Choose
n Z n
X 1 α 0
X
< σ >= C : dV = fα C : 0 .
V Vα
α=0 α=0
Hence
n n n
( )
X X X
C̄ ≥ 2 fα Cα : 1 (4s)
− α
fα C : fα Cα
α=0 α=0 α=0
where
Z
1
W (d ) = U (d )dV
2V V
Z
1
W0 (d ) = U0 (d )dV
2V V
Obviously, Πp () is convex and Πp () is concave.
Define stress polarization
∂U
pij = d (8.103)
∂ij
Subsequently, we can form the following Legendre-Fenchel transformation,
n o
Π∗p = sup < p : d > −Πp (d ) (8.104)
d ∈E
n o
Πp∗ = inf < p : d > −Πp (d ) (8.105)
d ∈E
where Z
d 1
< p : >= p : d dV
V V
and
1 ◦
2
E := ij ij ∈ L (V ), ij = (ui,j + uj,i ), and ui ∈ V
2
n o
V := ui ui ∈ L2 (V ), W (ui,j ), W0 (ui,j ) < ∞, ui = xj 0ij , ∀x ∈ ∂V
◦ n o
V := ui ui ∈ L2 (V ), W (ui,j ), W0 (ui,j ) < ∞, ui = 0, ∀x ∈ ∂V
Combining Eqs. (8.108) and (8.109), we have the original form of Talbot-
Willis variational princinple
inf {< p : d > +W0 (d )} − Π∗p (p)
{d ∈E}
≤ inf W (d ) ≤
{d ∈E}
where
Z
1 −1
Rπ , (orR̄π ) = −∆Cijk` pij pk` + pij dij + 2pij 0ij dV
2V V
8.8 Exercises
Probelm 8.1 Consider a functional
P : H 1 ([a, b]) → IR
where Z bq
P (u) = 1 + [u0 (x)]2 dx .
a
with essential boundary condition u(a) = ūa and u(b) = ūb .
228 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Find the first variation, second variation, and G^ateaux derivative. Derive
associated the Euler-Lagrange equation.
n · σ̄ = t0 (x), ∀x ∈ ∂V (8.113)
σij,j = 0, ∀x ∈ Ω (8.117)
nj σij = t0i , ∀x ∈ Γt , and Γu = ∅ (8.118)
1
ij = ui,j + uj,i (8.119)
2
∂Uc 1
ij = , Uc (σ) := Dijk` σij σk` . (8.120)
∂σij 2
0
Consider a comparison elastic solid with compliance tensor, Dijk` and
(0)
σij,j = 0, ∀x ∈ Ω (8.121)
(0)
nj σij = t0i , ∀x ∈ Γt , and Γu = ∅ (8.122)
(0) 1 (0) (0)
ij = ui,j + uj,i (8.123)
2
(0)
(0) ∂U0 (0) 1 (0) (0) (0)
ij = , U0 (σ) := D σ σ . (8.124)
∂σij
(0) 2 ijk` ij k`
Comparison Variational Principles 229
Let
(0)
d
σij = σij + σij (8.125)
(0)
ij = Dijk` σk` + qij (8.126)
Chapter 9
u = ū = x · ¯, ∀x ∈ ∂V (Γt = ∅)
u0 = ū = x · ¯, ∀x ∈ ∂V (Γt = ∅)
I¯
I ≤ inf W (d ) ≤ |{z} (9.1)
|{z} d ∈E
∆C>0 ∆C<0
Assume that there are n-phase in the composite (including the matrix). In
each phase (inclusion), the elastic tensor as well as stress polarization tensor is
constant, i.e.
n
X
C(x) = Cr H(Ωr ) (9.3)
r=1
n
X
p(x) = pr H(Ωr ) (9.4)
r=1
Bounds on Effective Properties 231
where H(·) is the Heaviside function, and Ωr is the domain of each phase,
1, ∀x ∈ Ωr
H(Ωr ) =
0, ∀x 6∈ Ωr
4
Z 1 Z n
1 0
X
p : dV = pdV : ¯ =< p >: ¯ = fr pr : ¯ (9.8)
V V V V r=1
5 Z n
1 d 1X
p : dV = − fr pr : Pr : pr − < p > (9.9)
2V V 2
r=1
where Z
P := r
Γ∞ (x0 − x)dVx0
Ωr
and
1 ∞
Γ∞
ijk` := − Gki,j` (x0 −x)+G∞
kj,i` (x0
−x)+G ∞
`i,jk (x0
−x)+G ∞
`j,ik (x0
−x)
4
232 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
How to integrate Z
1
p : d dV =? (9.10)
2V V
Consider the subsidiary condition,
(0)
Cijk` udk,`j + pij,j = 0 (9.11)
We solve udk in terms of pij by using Green’s function method. Conisder the
Green’s function of the comparison solid in an infinite medium, i.e.
(0)
Cijk` G∞ 0 0
km,`j + δim δ(x − x ) = 0, ∀ x, x ∈ IR
3
Multiplying Gim (x0 − x) with (9.11) and integrating it over V, one has
Z h i
(0)
Cijk` udk,` + pij G∞ 0
im (x − x)dVx0 = 0
V ,j
(0)
Let tij = Cijk` udk,` . Integration by parts yields,
Z h i
(0) d
G∞
im (x 0
− x) C u
ijk` k,` +p ij nj dS
∂V | {z }
tij
Z
∂ ∞ 0 h
(0) d
i
− G
0 im (x − x) C u
ijk` k,` + p ij dV
V ∂xj
Z Z h
∞ 0 ∂ ∞ 0 ih
(0) d
i
= Gim (x − x )[tij + pij ]nj dS − G
0 im (x − x) C u
ijk` k n ` dS
∂V ∂V ∂xj | {z }
=0
∂2
Z h Z
∞
ih
(0) d
i ∂ ∞ 0 0
+ 0 0 Gim Cijk` uk n` dV − 0 Gij (x − x)pij (x )dV
V ∂x j ∂x ` V ∂x j
Z Z
∂
= G∞ 0
im (x − x)[tij + pij ]nj dS −
∞ 0 0
0 Gij (x − x)pij (x )dV
∂V V ∂xj
Z
(0)
+ Cijk` G∞ 0 d 0
km,j` (x − x) ui (x )dV
V | {z }
−δim δ(x0 −x)
− G∞ 0
im,j (x − x) < pij > dV = 0 (9.13)
V
Thus substracting (9.13) from (9.12) will be affect the value of (9.12),
Z
0 0
d
um (x) = G∞ 0
im (x − x)[tij (x ) + (pij (x )− < pij >)]nj dS
∂V
Z
0
− G∞ 0
im,j (x − x)(pij (x )− < pij >)dV (9.14)
V
Now pij − < pij > also oscillates around zero, since its mean is zero, i.e.
< pij − < pij >>= 0. We can then neglect the boundary term, and finally we
have
Z
d
um (x) ≈ − G∞ 0 0
im,j (x − x)(pij (x )− < pij >)dVx0 (9.15)
V
Hence
Z h
1 i 0 0
dm` (x) = G∞
im,j` + G ∞
i`,jm (x − x) p ij (x )− < pij > dVx0 (9.16)
2 V
234 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
where
1h ∞ i
Γ∞
m`ij (y − x) := − Gim,j` + G∞ ∞ ∞
i`,jm + Gjm,i` + Gj`,im (y − x) (9.18)
4
(9.20)
Now we come back to evaluate (9.10). Let stress polarization p(x) is piece-
wise constant, i.e.
Xn
p(x) = pr H(Ωr )
r=1
n
X
<p> = fr pr
r=1
Therefore,
Z Z X n
1 1
p : d dV = pr H(Ωr ) :
2V V 2V V
r=1
Z 0
h 0
i
− Γ∞ (x − x) : p(x )− < p > dVx0 dVx
V0
Consider x ∈ Ωs . ps − < p > is constant inside Ωs . Thus,
Z
0
Γ∞ (x − x) : pr − < p > dVx0
V0
Z Z
∞ 0 0
= Γ (x − x)dVx0 + Γ∞ (x − x)dVx0 : pr − < p > dVx0
Ωs V 0 −Ωs
Assume that the RVE is a gigantic spherical ball and all Ωr are spherical inclu-
sions. By Mori-Tanaka lemma,
Z
0
Γ∞ (x − x)dVx0 = 0
V 0 −Ωs
In fact, for x ∈ Ωs
Z Z
∞ 0 0
Γ (x − x)dVx0 = Γ∞ (x − x)dVx0
V Ωs
because the integral over a spherical ball does not dependent on the size of
inlcusion (recall P = S : D).
Hence,
Z n n Z
1 d 1 XX
p : dV = − pr H(Ωr (x)
2V V 2V Ωr
r=1 s=1
Z
∞ 0
: Γ (x − x)dVx : ps H(Ωs (x) dVx
0
Ωs
n Z
1 X
+ pr H(Ωr (x)
2V Ωr
r=1
Z
∞ 0
: Γ (x − x)dVx0 :< p > dVx
Ωs
236 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Consider
1 r=s
H(Ωr (x))H(Ωs (x)) = (9.23)
0 r 6= s
and let Z
0
r
P := Γ∞ (x − x)dVx0
Ωr
We then have
Z n Z
1 d 1 X
p : dV = − dVx pr : Pr : pr
2V V 2V
r=1 Ωr
n Z
1 X
+ dVx pr : Pr :< p >
2V Ωr r=1
n
1X
= − fr pr : Pr : (pr − < p >)
2
r=1
n
1X
= − fr prij Pijk`
r
prk` − < pk` >
2
r=1
Pn r
where < pk` >= r=1 fr pk` .
Remark 9.1.1 Recall that by using Radon transform one can write,
Z
1 00
δ(x) = − 2 δ (ξn xn )dS
8π |ξ |=1
and consequently,
Z
1 −1
G∞
ij (x) = 2 Kij (ξ)δ(ξn xn )dS
8π |ξ |=1
where
1 + ν (0) 2(4 − 5ν (0) )
s1 = , s 2 =
3(1 − ν (0) ) 15(1 − ν (0) )
and for isotropic comparison solid,
1 1
D(0) = (0)
E(1) + E(2)
3K 2G(0)
Therefore,
(0)
s1 (1) s(0) (2)
P = E + E
3K (0) 2G(0)
1 + ν (0) (1) (4 − 5ν (0) )
= (0) (0)
E + (0) (0)
E(2)
9K (1 − ν ) 15G (1 − ν )
( )
1 1 (2) 2(4 − 5ν (0)
= − 1 ⊗ 1(2) + 1(4s) (9.24)
2G(0) (1 − ν (0) ) 15 15
3K (0) − 2G(0)
Consider ν (0) = . One can also have
2(3K (0) + G(0) )
K0 = K1 , K = K2 , and G0 = G1 , G = G2 .
Obviously that
¯ij = ¯δij
1
1
inf W (d ) = C̄ijk` ¯δij ¯δk`
d ∈E 2
1h (1) (2)
i
= 3K̄Eijk` + 2ḠEijk` (¯ )2 δij δk`
2
9 2
= K̄¯
2
(1) (2)
Note that Eijk` δij δk` = 3 and Eijk` δij δk` = 0.
2
1 (0)
W0 (0 ) = C (¯ δij )(¯
δk` )
2 ijk`
1h (1) (2)
i
= 3K1 Eijk` + 2G2 Eijk` (¯ )2 δij δk`
2
9
= K1 ¯2
2
3 Z
1
p : (0) dV = f1 p1 : ¯ + f2 p2 : ¯ = 3f2 p¯
V V
(1) (2)
4 Because pij = 0 and pij = pδij ,
Z 2
1 1X
p : ∆C−1 : pdV = fr pr : ∆C−1r : pr
2V V 2
r=1
1 f2 f2
= E(1) + E(2) p2 δij δk`
2 3(K2 − K1 ) 2(G2 − G1 )
f2 p 2
=
2(K2 − K1 )
Bounds on Effective Properties 239
¯ = 9 K2 ¯2 − f1 p2 1 f1 f2 p2 9 2
I(p) − 4 ≥ K̄¯
+ 3f1 p¯ (9.28)
2 2(K1 − K2 ) 2 K2 + 3 G2 2
¯
To find the maximum value of I(p), we examine the stationary condition,
∂ I¯ 3¯
= 0, ⇒ psta = (9.29)
∂p 1 f2
+
K1 − K2 K2 + 43 G2
240 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Figure 9.1. Variational Bounds for Bulk Modulus: (a) Medium one, and (b) Medium two.
Figure 9.2. Variational Bounds for Shear Modulus: (a) Medium one, and (b) Medium two.
(a) (b)
Figure 9.5. Examples of homogeneous isotropic (a) and homogeneous anisotropic media.
9.2.3 Applications
Example 9.1 Consider Voigt bound and Reuss bound,
X n Xn
fr Cr−1 ≤ C̄ ≤ fr Cr
r=1 r=1
Both these two bounds only require information of volume fraction of each
phase. Since volume fraction,
(r)
fr = S1 (x),
246 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Therefore,
Z Z X n
1 d 1
p : dV = pr I (r) (x) :
V V 2V V
r=1
Z h 0
i
− Γ∞ (x0 − x) : p(x − < p > dVx0 dVx
V0
n
Z X Z
1 00
= − pr I (r)
(x) : Γ∞ (x )
2V V r=1 V 00
n
hX n i
00
X
: ps I (s) (x + x ) − fs ps dVx00 dVx
s=1 s=1
n X
n Z
1 X 00
= − pr : Γ∞ (x )dVx00
2 V 00
r=1 s=1
Z
1 (r) (s) 00 (r)
: I (x)I (x + x ) − I fs ps dVx
V V
n n Z
1 XX 00
= − pr : Γ∞ (x )dVx00
2 V 00
r=1 s=1
00
(rs)
: S2 (x, x + x ) − fr fs ps
0
One the other hand, when kx − x k ≤ Rr or Rs . There can be only one
phase exists within the region, hence
(rs) 0 (r)
S2 (x, x ) = S1 δrs
To sum up 0
fr δrs kx − x k ≤ Rr
(rs) 0
S2 (x, x ) =
0
fr fs kx − x k > Rr
“When strolling along a sandy beach in shores most people choose the wet
strip left by retreating waves, which is hard and smooth enough to make the
walk more comfortable than the dry part of the beach. On the other hand, to
avoid their shoes and socks being soaked they must constantky watch the play
the surf licking the strip. This steady twisting of the neck becomes disagreeable
after a few minutes. There is, however, a remedy. Instead of looking sidewise
one keeps looking straight ahead; in every instant he sees the instantaneous
water edge and he directs his steps tangentially; he walks along a line touching
the edge in a single point without cutting contact lies far enough away to render
the variations small and easily accounted for: neither looking to the left, nor
sudden jumping to the right is necessary.
The background for the behavior I recommend here (after having tried it)
is the ‘ ergodic principle’: the distribution of water tongues licking the shore
in a fixed point observed during a long time is the same as the distributions
shown in a fixed moment by a long portion of the water edge — the principle
involved is the identity of time-distribution and space-distribution. To apply
it here the walker has to limit his observation to the part of the shore he will
cover in the next minute — in most cases such tactics keep him on the safe side
without leading him out of the wet strip of the beach. · · · · · · ”
I thought that some explanation may be needed to correctly understand
Steinhaus’ analogy:
What Steinhaus was trying to say is that consider an infinite set of good
weather day, if a person comes to a beach every afternoon at 2:00 clock he may
find that at a particular spot (fixed spatial location) the sea water line on the
beach is a stochastic event and all the measurement on water line on each day
consist of a statistic ensemble. We assume that there is a statistical average
value for the sea water line on that spot, which is the average in time. The
ergodic principle suggests that if a system is both homogeneous in space and
in time, one can then find that average without measuring water line at 2:00
pm on infinite days. Instead, he can just walk along a path that is tangential
to the water (shore) line on the beach, which is also assumed to “infinite”. By
Bounds on Effective Properties 249
doing so, the average position along his path on the beach may be equal to the
statistical average of the time ensemble.
Note that we do not consider the the surge or recede of sea water line due to
the effect of tide. Hence, the person who is in charge the measurement has to
come to the beach every afternoon at the same time (e.g. 2:00 pm), provided
that the weather is always good.
9.3 Exercises
Probelm 9.1 Show that for a spherical inlusion, Ω ⊂ V ,
Z
P := Γ∞ (y − x)dVy
Ω Z
1 ∞
= Γ̃ (ξ)dS (9.42)
4π |ξ |=1
Probelm 9.2 Consider a well-order two phase composite (K2 > K1 and
G2 > G1 ). Derive the Hashin-Shtrikman bounds for shear modulus,
f2 f1
G1 + ≤ Ḡ ≤ G2 +
1 6(K1 + 2G1 )f1 1 6(K2 + 2G2 )f2
+ +
G2 − G1 5(3K1 + 4G1 )G1 G1 − G2 5(3K2 + 4G2 )G2
(9.43)
Assume that K1 = 8GPa & G1 = 5GPa and K2 = 20.0GPa & G2 = 18GPa .
Plot the Voigt bound, Ruess bound, Mori-Tanaka, and Hashin-Shtrikman bounds
for both bulk modulus and shear modulus for comparison.
Hints:
Hashin, Z. and Shtrikman, S. [1961], “Note on a variational approach to
the theory of composite elastic materials,” The Frabklin Institute Laboratories,
pp. 336-341.
Hashin, Z. and Shtrikman, S. [1962a], “On some variational principles
in anisotropic and non-homogeneous elasticity,” Journal of Mechanics and
Physics of Solids, Vol. 10, pp. 335-342.
Hashin, Z. and Shtrikman, S. [1962b], “A variational approach to the the-
ory of the elastic behavior of polycrystals,” Journal of Mechanics and Physics
of Solids, Vol. 10, pp. 343-352.
Chapter 10
PERIODIC MICROSTRUCTURE
where |Y | is the volume of the unit cell. For a rectangular unit cell, |Y | =
8a1 a2 a3 .
Recall the definition of Fourier series in an 1D interval, [−a, a],
∞
X nπ
f (x) = F[f ](ξ) exp i x , n = 0, ±1, ±2, · · · ,
n=−∞
a
Z a
1 nπ
F[f ] = f (x) exp(−i x)dx
2a −a a
and the orthonormal condition
Z a
1
exp(ixξm ) exp(−ixξn )dx = δmn
2a −a
nπ mπ
where ξn = and ξm = .
a a
Accordingly, 3D orthonormal condition is
Z
1 1 ξ=ζ
exp(ix · ξ) exp(−ix · ζ)dVx =
|Y | Y 0 ξ 6= ζ
where ξ, ζ ∈ Λ, i.e.
nj π nk π
ξ = ξj ej = ej and ζ = ζk ek = ek .
aj ak
where
0 nj π
Λ = ξ = ξj ej ξj = , j = ±1, ±2, · · · , · · ·
aj
0
Note that the difference between index set Λ and Λ is that nj 6= 0, or ξ 6= 0.
When ξ = 0, Z
1
F[u](0) = u(x)dVx
|Y | Y
which is the average displacement field.
On the other hand, if the composite undergoes a rigid body translation,
u(x) = u0 , which is not periodic, one may find that
F[u](0) = u0
and Z
1
F[∇ ⊗ u](ξ) = ∇ ⊗ u(x) exp(−ix · ξ)dVx
|Y | Y
because Z
n · σ(x) exp(−ix · ξ)dS = 0
∂Y
by periodicity. In particular, when ξ = 0,
Z
n · σ(x)dS = 0
∂Y
= x · 0 , ∀x∂V
The stress fields in the matrix and in the second phase are
M M
σij = Cijk` (0ij + dij ), ∀x ∈ M = Y /Ω
Ω Ω
σij = Cijk` (0ij + dij ), ∀x ∈ Ω
ud+ d−
i = ui , ∀ x ∈ ∂Ω
Let,
X X
∗k` (x) = F[∗k` ](ξ) exp(iξ · x) = ˆ∗k` exp(iξ · x) (10.14)
0 0
ξ ∈Λ ξ ∈Λ
where
Z Z
1 1
ˆ∗k` = ∗k` exp(−iξ · x)dVx = ∗k` exp(−iξ · x)dVx
Y Y Y Ω
and
X
ui (x) = F[ui ](ξ) exp(ix · x) exp(iξ · x) = ûi (ξ) exp(iξ · x) (10.15)
0
ξ ∈Λ
258 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
where Z
1
ûi (ξ) = ui (x) exp(−iξ · x)dVx
|Y | Y
Note that uniform eigenstrain is excluded because it induces a divergent
displacement field, i.e.
Nik (ξ) M
F[ui ](ξ) := ûi (ξ) = −i C ∗ ξ` (10.17)
D(ξ) k`mn mn
Recall,
i X
dij (x) = ξi F[uj ](ξ) + F[udi ](ξ)ξj exp(iξ · x)
2
ξ ∈Λ0
1 CM ξ
`
gijmn (ξ) = ξi Njk (ξ) + ξj Nik (ξ) k`mn (10.18)
2 D(ξ)
CΩ : (0 + d ) = CM : (0 + d − ∗ ) .
We have
0 + d = (CM − CΩ )−1 : CM : ∗
and subsequently,
0 = AΩ : ∗ (x) − d (x)
This leads to the following integral equation,
∗
0ij − AΩ ijmn mn (x)
Z
X 1 0 0
+ gijmn (ξ) ∗mn (x ) exp(i(x − x ) · ξ)dVx0 = 0(10.20)
.
0
|Y | Ω
ξ ∈Λ
Z
1
This equation is difficult to solve. Calculate the average (10.20)dVx
|Ω| Ω
in the inclusion. One has
X 1 Z
0 = AΩ : ¯∗ − g(ξ) : exp(iξ · x)dVx
|Ω| Ω
∈Λ0
1 Z
0 0
· ∗ (x ) exp(−iξ · x )dVx0
|Y | Ω
Define a scalar function,
Z
1
g0 (ξ) = exp(iξ · x)dVx (10.21)
|Ω| Ω
M σ0 .
where ¯ij = Dijmn mn
The simplest approach to solve (10.22) is to replace ∗ (x) by its volume
average, i.e., ∗ (x) ≈ ¯∗ . Therefore,
X 1 Z 0
∗
0
= A : ¯ − Ω
g(ξ)g0 (ξ) exp(−iξ · x )dVx0 : ¯∗
|Y | Ω
ξ ∈Λ0
X
= AΩ : ¯∗ − f g0 (ξ)g0 (−ξ)g(ξ) : ¯∗
ξ ∈Λ0
X
= AΩ : ¯∗ − f G(ξ)g(ξ) : ¯∗
ξ ∈Λ0
Hence
1
G(ξ) = sin2 (ξa)
a2 ξ 2
where
q
η = a|ξ| = a ξ12 + ξ22 + ξ32
r
n1 π 2 n2 π 2 n3 π 2
= a + +
L L L
πa πa
q
= n21 + n22 + n23 = |n|
L L
Considering,
2 1/2 r
−1 2 1
J3/2 (η) = (η sin η − cos η) = (sin η − η cos η)
πη π η 3/2
and
9 h i2
G(ξ) = sin(a|ξ|) − a|ξ| cos(a|ξ|) .
a6 |ξ|6
One may find that for bcc precipitate cluster,
3
g0 (−ξ) = (sin η − η cos η) 1 + exp(−iξ · c)
η3
3
g0 (−ξ) = (sin η−η cos η) 1+exp(−iξ·c 1 )+exp(−iξ·c 2 )+exp(−iξ·c 3 )
η3
(a) (b)
Figure 10.4. Cluster of peripitates in various unit cells: (a) b.c.c. cluster, and (b) f.c.c cluster .
`
We define a small paramter = . Obviously, << 1. To separate the
L
effect of two scales, we introduce two coordinates: a fast coordinate and a slow
coordinate, which are defined as
y and x = y (10.26)
You may suggest that the slow coordinate is slowed by small parameter, . Or
vice versa,
x
x and y = (10.27)
1
You may suggest that the fast coordinate is speed up by a large paramter .
Then, the field variable u may be expressed in a two-scale representation:
u = u(x, y) By using chain rule, we may write
d ∂ ∂
= + (10.28)
dy ∂y ∂x
or vice versa,
d ∂ 1 ∂
= + (10.29)
dx ∂x ∂y
d du
E(y) = 0, 0 < y < L (10.30)
dy dy
Q(x, y0 ) = Q(x, y0 + `)
It leads to
Z y0 +`
dỹ
−y0 + D2 (x) = −(y0 + `) + D1 (x) + D2 (x) (10.38)
y0 E(ỹ)
Eq. (10.38) is called the solvability condition for inhomogeneous problem for
Q or u1 .
We then find that
1
D1 (x) = Z y0 +` (10.39)
1 dỹ
` y0 E(ỹ)
Periodic Microstructure 267
and hence Z y
dỹ
y0 E(ỹ)
Q(x, y) = −y + Z y0 +`
+ D2 (x) (10.40)
1 dỹ
` y0 E(ỹ)
Therefore,
y
Z
dỹ
y E(ỹ) ∂u
0
u1 (x, y) = −y + Z y00 +` + D2 (x) + ū1 (x)(10.41)
1 dỹ ∂x
` y0 E(ỹ)
∂u1 ∂u0 1 ∂u0
= − + 1 Z y0 +` dỹ ∂x (10.42)
∂y ∂x
E(y)
` y0 E(ỹ)
Next, we consider the differential equation at the second scale,
∂ h ∂u ∂u2 i ∂2u ∂u2 u1
1 0
2 : E(y) + + E(y) + =0. (10.43)
∂y ∂x ∂y ∂x2 ∂x∂y
Consider
∂ 2 u1 ∂ 2 u0 1 ∂ 2 u0
=− 2 + 1 y0 +` dỹ ∂x2
∂x∂y ∂x
Z
E(y)
` y0 E(ỹ)
Eq. (10.43) becomes
∂ h ∂u
1 ∂u2 i 1 ∂ 2 u0
E(y) + + Z y0 +` =0 (10.44)
|
∂y ∂x
{z
∂y
} 1 dỹ ∂x2
f unction of y ` y0 E(ỹ)
| {z }
f unction of x
Hence,
1 ∂ 2 u0
1 y0 +` dỹ ∂x2 = 0
Z
` y0 E(ỹ)
or
∂ 1 ∂u0
∂x 1 Z y0 +` dỹ ∂x = 0 . (10.45)
` y0 E(ỹ)
268 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
y0 +` Z 1 − f`
1 f ` dt
Z Z
1 dt 2 2 dt
= +
` y0 E(t) ` 0 E1 ` 0 E2
` − f` 1 f (1 − f )E2 + f E1
= + =
` E1 E2 E1 E2
and
1 E1 E2
Ee = y0 +`
= (10.47)
(1 − f )E2 + f E1
Z
1 dt
` y0 E(t)
The homogenized differential equation is,
d du0
Ee =0. (10.48)
dx dx
Periodic Microstructure 269
A u = f, ∀x ∈ Ω (10.49)
u = 0, ∀x ∈ ∂Ω (10.50)
where
∂ x ∂
A =− aij ( )
∂xi ∂xj
where x = (x1 , x2 , x3 ).
Define the fast coordinate,
x
y=
1
as if y is speed-up by the large paramter . We then can express the field
variable as a function of two independent scales, u (x) = u(x, y).
From chain rule, we have
∂ ∂ ∂ ∂yi ∂ 1 ∂
= + = +
∂xi ∂xi ∂yi ∂xi ∂xi ∂yi
We can then expand the differential operator, A , as
270 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
∂ 1 ∂ h ∂ 1 ∂ i
A = − + aij (y) +
∂xi ∂yi ∂xi ∂yi
h ∂ ∂ i h ∂ ∂ ∂ ∂ i
= −−2 aij − −1 aij (y) + aij (y)
∂xi ∂yj ∂xi ∂yj ∂yi ∂xi
h ∂ ∂ i
−0 aij (y)
∂xi ∂xi
= −2 A1 + −1 A2 + 0 A3 (10.51)
where
h ∂ ∂ i
A1 = − aij
∂xi ∂yj
h ∂ ∂ ∂ ∂ i
A2 = − aij (y) + aij (y)
∂xi ∂yj ∂yi ∂xi
h ∂ ∂ i
A3 = − aij (y)
∂xi ∂xi
Now we consider multiple scale expansion,
Hence
∂ ∂
A1 u = − aij (y) u
∂yi ∂yj
X ∂aij
= − ξj + iξi F[u](ξ) exp(iξy)
∂yi
ξ∈Λ
Therefore,
Z
1
F[F ] = 0 ⇒ F[F ](0) = F dVy = 0 .
|Y | Y
♣
To this end, we start to solve differential equations at each scale. At scale
−2 , we have
∂ ∂
A1 u 0 = − aij (y) u0 = 0
∂yi ∂yj
We claim that
u0 = u0 (x) .
That is the leading-order expansion is only the function of slow scale variable.
Since u0 is Y-peridoic, we have
X
u0 = F[u0 ](ξ) exp(iξy) .
ξ∈Λ
Consequently,
∂a
ij
X
A1 u 0 = 0 ⇒ − iξj + iξi F[u](ξ) exp(iξy) = 0 .
∂yi
ξ∈Λ
F[u](ξ) = 0 . (10.60)
Assume that
u0 = c(x)Q(y) + ū0 (x)
Eq. (10.60) becomes
Z
1
F[u](ξ) = c(x)Q(y) + ū0 (x) exp(−iξy)dVy
|Y | Y
Z
1
= c(x)Q(y) exp(−iξy)dVy = 0 (10.61)
|Y | Y
Z
because ū0 (x) exp(−iξy)dVy = 0 when ξ 6= 0.
Y
The only possibility that (10.61) holds is that Q(y) = 1 or Q(y) = 0. In
either case, u0 = u0 (x). We proved our claim.
Next, we consider the differential equation at scale −1 :
A1 u 1 + A2 u 0 = 0 .
Periodic Microstructure 273
Hence
∂aij ∂u0
A1 u 1 = (10.62)
∂yi ∂xj
This suggets the following separation of variable,
∂u0
u1 (x, y) = Uk (y) + ū1 (x) (10.63)
∂xk
and subsequently,
∂u
0
A1 u 1 = A1 Uk (y)
∂xk
∂ ∂U ∂u
k 0
= − aij (y) (10.64)
∂yi ∂yj ∂xk
Combining (10.62) and (10.64), we find the canonical equation for a unit cell
problem,
∂aik ∂ ∂U
k
+ aij (y) =0. (10.65)
∂xi ∂yi ∂yj
with the possible boundary conditions at interface of different phases,
h i h ∂Uk i
Uk = 0, and aik + aij ni = 0 (10.66)
∂xj
A1 u 2 + A2 u 1 + A3 u 0 = f
The condition that equation (10.67) has a unique periodic solution is that
That is Z
1
A2 u1 + A3 u0 dy = f (10.68)
|Y | Y
274 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Consider
u0 = u0 (x)
∂u0
u1 = Uj + ū1 (x)
yj
One can show that
∂ 2 u0
A3 u0 = −aij (10.69)
∂xi ∂xj
h ∂ ∂ ∂ ∂ i ∂u0
A2 u1 = − aij (y) + aij (y) Uk (y) + ū1
∂xi ∂yj ∂yi ∂xj ∂xk
∂Uk ∂ u0 2 ∂ 2
∂ u0
= −aij − aij (y)Uk (y)
∂yj ∂xi ∂xk ∂yi ∂xj ∂xk
∂ ∂ ū
1
− aij (y) (10.70)
∂yi ∂xj
Change the dummy indices j ↔ k in the first term of (10.70). We can write
that
∂Uj ∂ 2 u0 ∂ ∂ 2 u0
A2 u1 + A3 u0 = − aij + aik − aij (y)Uk (y)
∂xk ∂xi ∂xj ∂yi ∂xj ∂xk
∂ ∂ ū1
− aij (y)
∂yi ∂xj
Via divergence theorem,
∂ 2 u0
Z Z
1 1 ∂Uj
(A2 u1 + A3 u0 )dy = − aij + aik dVy
|Y | Y |Y | Y ∂xk ∂xi ∂xj
h i h i
− (aij (y)Uk (y)u0,jk (x) ni − aij (y)ū1,j ni
AH u0 = 0, ∀x ∈ Ω (10.73)
u0 = 0, ∀x ∈ ∂Ω (10.74)
∂T0 (x)
T1 (x, y1 ) = Uα (y1 ) , α = 1, 2
∂xα
Note that first we assume that the mean temperature at this scale is zero, i.e.
T̄1 (x) = 0; and Uα (y1 ) are Y-periodic functions that are the following 1D
canonical cell problem,
Integrate (10.78),
dUα (y1 )
−λ11 (y1 ) = λ1α (y1 ) − Cα
dy1
dUα (y1 ) λ1α (y1 ) Cα
⇒ =− +
dy1 λ11 (y1 ) λ11 (y1 )
In specific,
Z ` −1
C1 = λ−1
11 (ξ)dξ
0
Z `
λ12 (ξ)λ−1
11 (ξ)dξ
0
C2 = Z `
λ−1
11 (ξ)dξ
0
Consequently, we find the closed form solution for canonical cell problem,
Z y1
λ−1
11 (ξ)dξ
U1 (y1 ) = −y1 + Z0 ` (10.81)
λ−1
11 (ξ)dξ
0
Z y1
U2 (y1 ) = − λ12 (ξ)λ−1
11 (ξ)dξ
0
Z `
λ12 (ξ)λ−1
11 (ξ)dξ Z y1
0 −1
+ Z ` λ 11 (ξ)dξ (10.82)
0
λ−1
11 (ξ)dξ
0
Z
1 ∂Uj
λ̄ij := aij + aik dy .
|Y | Y ∂xk
1 `
Z
∂U1
λ̄11 = λ11 (ξ) + λ11 (ξ) (ξ) dξ
` 0 ∂y1
1 `
Z 1
= λ11 − λ11 + C1 dy = C1
` 0 `
Z `
1 −1
= ( λ−1 (ξ)dξ
` 0 11
278 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
and
1 `
Z
∂U2
λ̄12 = λ12 (ξ) + λ11 (ξ) (ξ) dξ
` 0 ∂y1
1 `
Z
= λ12 (ξ) − λ12 (ξ) + C2 dξ
` 0
1 `
Z
λ12 (ξ)λ−1
11 (ξ)dξ
` 0
= = λ̄21
1 ` −1
Z
λ (ξ)dξ
` 0 11
and
1 `
Z
∂U2
λ̄22 = λ22 (ξ) + λ21 (ξ) (ξ) dξ (10.83)
` 0 ∂y1
1 `
Z
= λ22 (ξ) − λ212 λ−1
11 (ξ) + C 2 λ 12 λ −1
11 (ξ) dξ (10.84)
` 0
1 ` 1 `
Z Z
= λ22 (ξ)dξ − λ12 (ξ)λ−1
11 (ξ)dξ
` 0 ` 0
Z ` 2
λ12 (ξ)λ−1
11 (ξ)dξ
1 0
+ Z ` (10.85)
`
λ−1
11 (ξ)dξ
0
such that
āij ξi ξj = min J U (10.93)
1 (Y )
U∈H#
280 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
functions, i.e.
n o
1
H# (Y ) := u u is Y − periodic, and u ∈ H 1 (Y )
that is Z
(u2 + |∇u|2 )dy < +∞
Y
To show this, we first show that the Euler-Lagrange equation of J(U) is the
governing equation of canonical cell problem.
Assume that aij is symmetric and real. It subsequently implies that aij is
positive definite. Therefore,
Z
1 ∂δU ξ
k k ∂U` ξ` ∂Uk ξk ∂δU` ξ`
δJ = aij (y) ξj + + ξi + dy
|Y | Y ∂yi ∂yj ∂yi ∂yj
Z
2 ∂Uk ξk
= aij (y) ξi + δU` ξ` dS
|Y | ∂Y ∂yi
Z
2 ∂ ∂Uk ξk
− aij (y) ξi + δU` ξ` dy = 0
|Y | Y ∂yi ∂yi
By periodic conditions
Z
2 ∂Uk ξk
aij (y) ξi + δU` ξ` dS = 0,
|Y | ∂Y ∂yi
it then leads to
Z
2 ∂ ∂Uk
δJ = − aij (y) δik + δU` ξk ξ` dy = 0
|Y | Y ∂yi ∂yi
and hence
∂ ∂Uk
− aij (y) δik + δU` = 0 .
∂yi ∂yi
Consider Uk = 0 ∈ H# 1 (Y ). One can find an upper bound for effective
1 In music, the sign # is used to indicate that a note is to be raised by a half tone. Similar meaning implies
It is obvious that
āij ξi ξj ≥ min Jc (ζ) (10.96)
ζ ∈L2# (Y ) and
Y ζ (y)dy=0
R
where
Z
1
Jc (ζ) := aij (ξi + ζi (y))(ξj + ζj (y))dy
|Y | Y
Z
−2Ck ζk (y)dy − 0 (10.97)
Y
Cj =< a−1
ji (y) >
−1
ξi (10.102)
282 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Y ζ (y)dy=0
R
Z
1
= Cj (ξi + ζi )dy = Cj ξj
|Y | Y
= < a−1 −1
ji >Y ξi ξj
1 Z −1
= a−1
ij (y)dy ξi ξj
|Y | Y
From the above estimate, we find a lower bound for effective coefficient, āij ,
i.e.
1 Z −1
āij ≥ a−1 (y)dy . (10.103)
|Y | Y ij
which is the so-called Reuss bound.
Hence
∂ ∂ 1 ∂ 0
uk,` = uk = + uk + u1k + 2 u2k + · · ·
∂x` ∂x` ∂y`
= −1 eY k` (u(0) ) + 0 (eXk` (u(0) + eY k` (u(1) )) +
+1 (eXk` (u(1) ) + eY k` (u2 )) + · · · (10.109)
and
σij (x, y) = Cijk` (y)uk,`
h
(0) (1) (2)
= Cijk` (y) −1 uY k,` + 0 (u0Xk,` + u1Y k,` ) + (uXk,` + uY k,` )
i
+···
(0) (1) (2)
= −1 σij + 0 σij + 1 σij + · · · (10.110)
and
(0)
0 : ui = ūi ;
(1)
1 : ui = 0; ∀x ∈ Γu (10.118)
(2)
2 : ui = 0;
······
We first examine the leading order equilibrium equation and boundary con-
dition,
(0)
∂σij
=0
∂yj
This yields
(0) (0)
σij = σij (x)
On the other hand
(0)
(0) ∂u
σij = Cijk` (y) k
∂y`
To commodate both conditions, we have to set
(0)
σij = 0 . (10.119)
and
(0) (0)
ui = ui (x) (10.120)
To solve the second order boundary-value problem, the follwing separation
of variable is adopted
(0)
(1) k` ∂uk (1)
ui (x, y) = χi (y) (x) + ūi (x) (10.121)
∂x`
Periodic Microstructure 285
where the unknown vector function, χk` i (y)ei , is often referred to as the char-
acteristic displacement field. We further assume that
(1) k` ∂u0k
σij (x, y) = σ̂ij (y) (x) (10.122)
∂x`
Consider
(1) (0)
∂ui ∂χk`
i ∂uk
= (10.123)
∂yj ∂yj ∂x`
and
(1) (0) (1) (0) (1)
σij = Cijk` uXk,` + uY k,` = Cijk` eXk` + uY k,` . (10.124)
We find that
(1)
mn ∂χmn
k
(0)
σij = Cijk` Tk` + uXm,n (10.125)
∂y`
mn = 1 δ
where Tk` mn (0) (0)
km δ`n + δkn δ`m , because Tk` uXm,n = eXk` .
2
Accordingly,
mn
mn ∂χmn
k
σ̂ij = Cijk` Tk` +
∂y`
Then the equilibrium equation on second scale (−1 ) provides the governing
equation for the canonical cell problem,
(1) mn (0) mn
∂σij ∂ σ̂ij ∂um ∂ σ̂ij
= 0, ⇒ = 0, ⇒ =0. (10.126)
∂yj ∂yj ∂xn ∂yj
More explicitely, the governing equation for canonical cell problem is
∂ h
mn ∂χmn
k
i
Cijk` Tk` + = 0, ∀y ∈ Y
∂yj ∂y` (10.127)
The related interface continuity conditions and periodic conditions are omited
here.
Consider the equilibrium equation at third scale (0 ). We have
(2) (1)
∂σij ∂σij
= − fi + = Fi , ∀ y ∈ Y
∂yj ∂xj
The Fredholm alternative condition requires that
Z
1
Fi (y)dy = 0 .
|Y | Y
286 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
mn 1 ∂P`mn ∂Pkmn mn
Tk` = + = P(k,`)
2 ∂yk ∂y`
Therefore, the weak formulation for the canonical cell problem can be written
as Z
1
mn
Cijk` (y) P(k,`) + χmn
(k,`) v(i,j) dVy = 0 . (10.135)
|Y | Y
Define the bilinear form
Z
1
aY (u, v) = Cijk` (y)u(i,j) v(k,`) dVy (10.136)
|Y | Y
aY (Pmn + χmn , v) = 0, ∀ v ∈ H#
1
(Y ) (10.137)
Once χmn
k,` being determined, the effective elastic stiffness tensor can then
be calculated based on definition
Z
H 1 k`
Cstk` = Cstmn (y)(Pm,n + χk`
m,n (y))dVy (10.138)
|Y | Y
288 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
H ij H 1
H
Cstk` Tst = Cstk` δsi δtj + δsj δti == Cijk` (10.139)
2
and
Z
H 1
k` ij
Cijk` = Cstmn Pm,n + χk`
m,n )Tst dVy
|Y | Y
Z
1
k` ij
= Cstmn Pm,n + χk`
m,n )P(s,t) dVy
|Y | Y
= aY (Pk` + χk` , Pij dVy (10.140)
is:
Find u(0) (x) ∈ VΩ such that
Z Z Z
H (0)
Cijk` u(k,`) v(i,j) dVx = fi vi dVx + t0i vi dS, ∀ v ∈ VΩ . (10.146)
Ω Ω Γt
where v = vi ei .
1 Solve the canonical cell problem on Y first, i.e. find χk` (y) ∈ H#
1 (Y ) by
solving
aY Pk` + χk` , v = 0, ∀v ∈ H# 1
(Y )
Note that the physical meaning of L∞ (Ω) space is that its occupant func-
tions satisfying the condition |u(x)| < ∞ almost everywhere in Ω.
We use the short-handed notation, → 0 to denote a limit process of a
sequence = {1 , 2 , · · · n , · · · · · · }, and n → 0 as n → ∞.
The strong convergence of a function sequence, u := {u1 , u2 , · · · , un , · · · },
is measured by the distance in the particular normed space, i.e. a sequence, u ,
is said to converge strongly in Lp (Ω) to a limit u0 , if
u → u0 , in Lp (Ω) strongly
u * u0 in Lp (Ω) weakly .
Periodic Microstructure 291
u (x) → u0 (x)
2 Assume that the sequence u (x) is bounded in Lp (Ω) (1 < p ≤ ∞), and
Then
u (x) → u0 (x) in Lq (Ω) (1 ≤ q < p) strongly .
It is worth noting that the product of two strong convergence sequence does
converge to the product of the two limits strongly, but it may be in a different
Lebesgue space in general.
Periodic Microstructure 293
ku v − u0 v0 kL2 (Ω) = k(u − u0 )(v − v0 ) + (u − u0 )v0 + (v − v0 )u0 kL2 (Ω)
1/2 1/2
≤ ku − u0 kL2 (Ω) kv − v0 kL2 (Ω)
1/2
1/2
+kv0 kL2 (Ω) ku − u0 kL2 (Ω)
1/2
1/2
+ku0 kL2 (Ω) kv − v0 kL2 (Ω)
Hence
u v → u0 v0 in L2 (Ω) strongly .
Unfortunately, the same is not true for the weakly convergent sequences. In
our previous example,
x
u (x) = sin → 0 in L2 (Ω) weakly
but for u (x) = v (x) = sin x
1
u (x)v (x) * ! in Lp (Ω) 1 ≤ p < +∞ .
2
Moreover, in practice, if u * u0 in Lp (Ω), and J(u) is a nonlinear func-
tional, say quadratic functional, J : Lp (Ω) → IR.
It is usually
J(u ) 6* J(u0 ) in any sense !
10.6.2 G- Convergence
Consider our model homogenization BVP,
x
L u = f, x ∈ Ω, where L = −∇ · A( ) · ∇
u = ū, ∀x ∈ ∂Ω
∂Ω
where the heat conduction (or diffusion) coefficient Aij (y) are Y-periodic func-
tions.
Suppose that solution of the above BVP can be found as
−1
u (x) = L f,
Suppose that there exists a weak limit u0 (x) in H 1 (Ω) such that
u (x) * u0 (x) in H 1 (Ω) weakly
and the weak limit u0 (x) has the representation,
Z −1
u0 (x) = G0 (x − y)f (y)dy =: L0 f
Ω
Let Ω be a bounded open set in IRd and define the space L∞ (Ω; Msα,β ) of
admissible symmetric coefficient matrices.
We have the following definition of G-convergence,
−∇ · A ∇u = f, x ∈ Ω
u = ū, ∀x ∈ ∂Ω
−∇ · A0 · ∇u0 = f, x ∈ Ω
u0 = ū, ∀x ∈ ∂Ω
Example 10.9 In this example, suppose that we have two objects with the
same macroscopic dimensions but different checkerborad microscopic struc-
ture.
The diffusitity matrix coefficients are assumed to be
Aij = aδij
We denote the diffusitivity in the white region as a1 and the diffusitivity in the
black region as a2 , and a2 > a1 .
We denote the first micro-structure as S1 and the second micro-structure as
S2 .
Obviously, the first sequence A (S1 ) and the second sequence A (S2 ) have
the same G-limit, i.e.
a0 (S1 ) = a0 (S2 ) .
As one can see that there is no pointwise convergence possibility, because for
a fixed spatial point,
Nevertheless, in this example, indeed, the weak convergence limit of the two
layouts are the same
Example 10.10 In this second example, we would like to show a case that
there are two micro-structure layouts with the same weak convergence limits,
but different G-limits.
In this example, we assume that in each unit cell, the black and white areas
are the same, therefore the volume fraction of the two phases are the same.
In the layout A, all the “good” material are connected, therefore it is a bet-
ter arrangement for heat conduction, whereas in the layout B, all the “good”
materials are isolated, disconnected, or insulated, it should be very hard for
heat to diffuse from one point to another point.
Based on this argument, the two layouts should have different G-limit, and
as indicated above.
Example 10.11 In the third example, we would like to show a case in which
two microstructure layouts have the same G-limit but different weak conver-
gence limits.
In this example, we fix the second layout of the previous example. Therefore,
we know that the G-limit of the second layout will be bounded by Vogit upper
bound and Reuss lower bound, i.e.
2a1 a2 1
Reuss bound = ≤ a0 S2 ≤ (a1 + a2 )
a1 + a2 2
298 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
We know change the first layout by increase the volume fraction of insolated
white phase, f1 such that f1 ∈ [0.5, ) and f1 → 1. Therefore, the G-limit of
the first layout will be bounded by
1
≤ a0 (S1 ) ≤ f1 a1 + (1 − f1 )a2 (10.150)
f1 1 − f1
+
a1 a2
Initially when f1 = 0.5 we have,
2a1 a2 1
< a0 S2 < a0 S1 < (a1 + a2 )
a1 + a2 2
If a1 << a2 .
The Reuss bound for the second layout is almost ≈ 2a1 . From Eq. (10.150),
one can see that as f1 → 1, the Reuss bound (lower bound) of the first layout
will become
1
→ a1 , as f1 → 1 .
f1 1 − f1
+
a1 a2
This suggests that at certain volume fraction, 0.5 < f1 = fw < 1.0, the
G-limits of the two layouts will be the same, i.e.
a0 (S1 ) = a0 (S2 ) .
At that moment, since fw > 0.5 6= f2 , the weak convergence limits of the
two layouts will not be the same, i.e.
10.6.3 H- Convergence
H-convergence is a generalization of G-convergence, in which, the differ-
ential operator A , or its coefficient matrix, does not require to be symmetric
anymore.
satisfies
u (x) → u0 (x) weakly in H1 (Ω) (10.153)
h iN
A · ∇u → a∗ · ∇u0 weakly in L2 (Ω) (10.154)
10.6.4 Γ- Convergence
For a large class of elliptical BVPs, each BVP under consideration has
one-to-one correspondence to a variational principle. The well-known Lax-
Milgram theorem guarantees the equivalence between the two.
Therefore, the convergence of differential operators may imply a possible
convergence of the corresponding functional in the related function spaces.
Definition 10.13 (Definition of Γ-Convergence) Let X be a func-
tional space endowed with a norm k·kd . Let be a sequence of positive indexes
which goes to zero. Let F be a sequence of functional defined on X with val-
ues in IR. The sequence F is said to Γ-convergence to a limit functional F0 if,
for any function x ∈ X,
1 all sequences x converging to x satisfy
F0 (x) ≤ lim inf F (x )
→0 x∈X
and
2 there exists at least one sequence x converging to x, such that
F0 (x) = lim F (x )
→0
10.7 Exercises
Probelm 10.1 Show that for isotropic materials the fourth-order tensor,
1 h i
gijk` (ξ) = ξj (δ i` ξk + δ ik ξ` ) + ξi (δ j` ξk + δ jk ξ` )
2ξ 2
1 ξi ξj ξk ξ` ν ξi ξj i
− + δk` . (10.159)
1 − ν ξ4 1 − ν ξ2
Probelm 10.2 Consider cuboidal region of inelastic strain (eigenstrain)
due to solute segregation forming cuboidal precipitates. The precipitate sub-
domain (or inclusion) has the dimension 2a×2a×2a, and the unit cell (U) has
the dimension 2L × 2L × 2L. The eigenstrain is assumed to have a constant
value within each inclusion, and be zero outside the inclusion,
δij ε, ∀x ∈ Ω;
ε∗ij = (10.160)
0, ∀x ∈ U/Ω,
where
n o
U x −L ≤ xi ≤ L, i = 1, 2, 3
= (10.161)
n o
Ω = x −a ≤ xi ≤ a, i = 1, 2, 3 , and a < L (10.162)
Find :
(a) the disturbed displacement field u1 (x) (Hint: Mura’s book pages: 20-
21).
(b) G(ξ) = g0 (ξ)g0 (−ξ).
Probelm 10.3 Consider the followin boundary-value problem in a medium
with periodic structure,
∂ 2 u
− = f, ∀x ∈ Ω (10.163)
∂xi ∂xi
u = 0, ∀x ∈ ∂Ω (10.164)
∂u
= 0, ∀x ∈ Γ (10.165)
∂n
where Γ is the interface between the matrix and inhomogeneous phase.
Show that the homogenized differential equation is
∂ 2 u0
−qik = f, ∀x ∈ Ω
∂xk ∂xk
Periodic Microstructure 301
him in the same office for almost four years (I was a postdoctal fellow then and
he was an emeritus professor).
Almost every week, he took me to lunch (because he insisted to pay ev-
erytimes, so we can not go out everyday), and I learned a lot of things from
Professor Mura, and had many good conversations as well as good memories.
Last year, Professor Mura received the Japanese Imperial model—the highest
honor bestow by Janpanes emperor and Royal family to scientists and other
citizens—for his contribution in micromechanics. I remembered back in 1997,
in his retirement party, professor Jan Achenbach said that Professor Mura is
one of the “seven samurai” (an international renowed Japanese moive, samurai
in Japanese means warrior, previously in Northwstern there were seven fa-
mous Mechanics professors: Achenbach, Belytschko, Dundurs, Keer, Mura,
Nemat-Nasser, and Bazant). Professor Mura is a theoretician, and has a very
“romantic” outlook of the world, (romantic is opposed to the “down-to-earth”
mentality of experimentalist) he believes that you are at your most creative
stage, when you are in your dream.
Biography sketch of Toshio Mura.
“ Toshio Mura, second son of Shinzo and Chie Fujii, was born in Ono, a
small port village of Kanazawa, the capital of Ishikawa Prefecture, Japan, on
December 7, 1925. Among the locals, the Fujiis are well known as brewers
having a long history in the area. Kanazawa is an old city on the coast of the
Periodic Microstructure 303
also noteworthy for introducing the concept of a dislocation flux tensor, which
is yseful when the dynamic motion of dislocations is examined. The period,
during which Mura’s Formula was found, coincided with his promotion to As-
sociate Professor of Civil Engineering.
....
The dislocation density and flux tensors were applied to continuum plastic-
ity theory. Believing that a stress appearing within the framework of continuum
plasticity was the sum of external and dislocation stresses, Mura published a
series papers, in the late 1960s, along these lines that emphasized the distribu-
tion and stress of dislocation.
In 1967 Mura became Professor of Civil Engineering. At that time Mura
nad J. G. Kunag, his student, obtained the solutions for a pile-up of edge dis-
locations against the interfacial boundary between different materials.
The pioneering work of J. D. Eshelby, his beloved peer, appears to have
inspired and stimulated Mura, as seen in his studies of static and dynamic
fields of dislocations in anisotropic media and in dislocation pile-ups. As can
be inferred from the preface to his book, Micromechanics of Defectcs in solids,
Mura regards Eshelby’s work on inclusions and inhomogeneities as being the
most important and fundamental.
To Mura the evaluation of the disturbance in elastic fields due to elastic in-
homogeneities is the most interesting application of the theory of inclusions.
For example, Z. A. Moschovides and Mura solved the stress field caused by
two inhomogeneities by applying the equivalent inclusion method with poly-
nomial eigenstrains. A computer program, performing the numerical calcula-
tions, complained that the matrice involved for linear equations were singular.
Moschovides looked for the bugs that might have caused this complaint, but no
bugs were found. The linear equations were carefully examined analytically
and the cause of the complaint was found. There existed certain distributions
of eigenstrains that yields no elastic field. Rozo Furuhashi, a visiting scholar,
and Mura later generalized this finding and showed that impotent inclusions
exist in a general sense. The impotent inclusions have eigenstrains defined by
derivatives of a continuous vector (displacement) that vanished at the bound-
ary of the inclusions. This anecdote illustrates Mura’s teachings: "study and
examine a specific subject carefully. If there is anything strange and exciting,
you can later generalize it in a broader sense.”
Mura also interacted with experimentalists, who eagerly sought his advice
and aid on issues of mathematics and mechanics. In particular, Morris E. Fine,
and his students in Northwestern’s Department of Materials Science and Engi-
neering, benefited from this interaction in their studies of the fatigue of alloys.
Mura also gained insight into material properties and structures by the interac-
tions with these materials scientists.
........
Periodic Microstructure 305
Chapter 11
−1
Note that ˙∞ ∞
ij = Dijk` σk` and Dijk` = Cijk` .
Since inside the void, there is no stress σij = 0, we have
˙ij = ˙∞ d ∗
ij + ˙k` = ˙ij
This means that eigenstrain rate should be the same as the actual strain rate,
which gives the physical meaning for eigenstrain rates. That is the prescribed
eigenstrain rate should be the expansion rate of the void.
308 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Therefore,
σ = C : (1(4s) − S) : ˙ .
Denote
Q := C : (1(4s) − S) .
The remote stress can be related with volumetric strain rate of the void, i.e.
Consider,
Consequently, we have
8 η(1 + ν)
T = ˙
3 (1 − ν)
Since the volume of the void is,
4π 3
V = a ⇒ V̇ = 4πa2 ȧ,
3
The relative void growth rate will be
V̇ ȧ
= 3 = 3˙
V a
Micromechanics Theory of Void Growth 309
8 η(1 + ν) V̇
T =
9 1−ν V
The above solution was obtained by Budiansky et al in 1981, almost ten
years after publication of the McClintock solution and the Gurson model.
ti = σij nj = Σij nj ∀x ∈ ∂V
where Σij is a contant tensor, and it is often denoted as the macro-stress tensor.
The following averagy theorems hold in the RVE,
To prove the additional strain rate formula, we use the so-called reciprocal
theorem of virtual power. Consider two sets of traction boundary conditions
and the corresponding velocity fields on the same ilinear viscous RVE, V , the
following equality holds,
Z Z
(1) (2) (2) (1)
t i u̇ i dS = ti u̇i dS
∂Ω− ∂Ω−
S S
∂V ∂V
where δ Ė = D : δΣ.
Let the traction b.c. for the second state as
(2) n · Σ, ∀x ∈ ∂V
t =
0, ∀x ∈ ∂Ω−
1
1
n · δΣ · u̇ = δΣ : (u̇ ⊗ n + n ⊗ u̇)
2
2
n · Σ · x · δ Ė = δΣ : D : (x ⊗ n) · Σ
We then have
Z Z Z
1
δΣ : { D : (x⊗n)·ΣdS − n⊗ u̇dS − n⊗ u̇dS} = 0 . (11.4)
V ∂V ∂V ∂Ω
Consider
1 1 Z
D: x ⊗ ndS · Σ = Ė; (11.5)
V ∂V
2 Z Z
1 1
Sym n ⊗ u̇dS = ˙
dV =< ˙ > (11.6)
V ∂V V V
3
Z Z
1 1
Sym n ⊗ u̇dS = − n ⊗ u̇dS
V ∂Ω− V ∂Ω
Z
1
= − n ⊗ u̇ + u̇ ⊗ n dS (11.7)
2V ∂Ω
Hence,
du̇
˙r = (11.12)
dr
u̇
˙θ = (11.13)
r
The incompressible condition yields,
du̇ u̇
˙r + ˙θ + ˙z = + + ˙z = 0 .
dr r
314 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Consequently,
˙
z
ru̇(r) − bḃ = − r2 − b2
2
˙z 2
⇒ ru̇ = bḃ + r − b2
2
Finally
b2 ḃ ˙z ˙z r
u̇(r) = + − (11.14)
r b 2 2
Let,
1
σ = (σr + σθ + σz ) .
3
We have
sr = σr − σ (11.15)
sθ = σθ − σ (11.16)
The components of the flow rule in an axisymmetric plane are
p p
3 ¯˙ 3 ¯˙
˙r = ˙pr = sr = (σr − σ) (11.17)
2 Y 2 Y
p p
3
¯˙ 3 ¯˙
˙θ = ˙pθ = sθ = (σθ − σ) (11.18)
2 Y 2 Y
(11.17) - (11.18) leads to
p
3 ¯˙
˙θ − ˙r = (σθ − σr ) (11.19)
2 Y
Utilizing (11.19), it can be found that
σθ − σr 2Y ˙θ − ˙r
=
r 3r ¯˙p
Micromechanics Theory of Void Growth 315
Subsquently,
¯˙p = ˙x (1 + x2 )1/2 (11.27)
and
b2 ḃ ˙z √ 2 b2 ḃ 1 √
˙θ − ˙r = 2 + = 3˙ z √ + = 3˙z x (11.28)
r2 b 2 3 r2 b˙z 2
Since
2b2 2 ḃ 1 2
dx = − 3
√ + dr = − xdr,
r 3 b˙z 2 r
dr 1 dx
=− .
r 2 x
Make change of variable,
b2
x=α ,
r2
and
r = b, x → α; r → ∞, x → 0 .
We can then integrate (11.22)
√
2 ∞ (˙θ − ˙r ) dρ 2 ∞
Z Z
σ∞ 3˙ x dr
= p = √ z
Y 3 b ˙
¯ ρ 3 b ˙z 1 + x2 r
Z 0 Z α
1 dx 1 dx
= −√ √ =√ √
3 α 1+x 2 3 0 1 + x2
1 α 1
= √ arcsinhx = √ arcsinh(α)
3 0 3
The inverse expression of the above result is
2 ḃ 1 h √3σ i
∞
√ + = sinh (11.29)
3 b˙z 2 Y
Based on uniaxial tension test, one can measure
Y
q
0
τ0 = J2 = √
3
We obtain the relationship between void growth rate and remote stress value,
√
ḃ 3 hσ i 1
∞
= ˙z sinh − ˙z (11.30)
b 2 τ0 2
A few comments about the McClintock solution are as follows:
Micromechanics Theory of Void Growth 317
1 McClintock solution is the only (essential) exact solution available for void
growth in nonlinear viscous media;
2 McCintock solution reveals an exponential increase in the void growth rate
under the positive remote stress load.
To illustrate the fact, we consider a finite cylindrical void with a heigh, H,
and radius b. The volume of the cylinder is
Thereby,
Ω̇ ḃ
= 2 + ˙z
Ω b
and hence √
Ω̇ 3 h √3σ i
∞
= ˙z sinh (11.31)
Ω 2 Y
Compare (11.31) with Budiansky et al’s linear viscous void solution,
Ω̇ 9 1−ν
= σ∞
Ω 8 η(1 + ν)
One may appreciate the significant difference between the two.
3 At the remote boundary, x ∈ ∂V ,
1
˙ ˙r = ˙θ = − ˙
˙z = ,
2
Hence the macro equivalent strain rate is
h2 i1/2 h21 1 i1/2
˙∞
eq = ˙∞ ∞
ij ˙ij = ˙2 + ˙2 + ˙2 = ˙ (11.32)
3 3 4 4
Ω = πa2 H
The void growth rate and relative void growth rate are
dΩ
= 2πaȧH + πa2 Ḣ (11.40)
dt
Ω̇ ȧ Ḣ
= 2 + (11.41)
Ω a H
Since,
ȧ Ḣ
= ˙rr (a) and = Ė33 ,
a H
one may find that
Ω̇ Ė
33 A 2A
=2 − − 2 + Ė33 = − 2
Ω 2 a a
which leads to
a2 Ω̇
A=− . (11.42)
2 Ω
That is: A is proportional to the relative void growth rate.
Since the matrix is a rigid-perfectly plastic von-Mises material, it obeys the
following flow rule,
2 σy
sij = ˙ij
3 ˙eq
where the effective strain rate can be explicitely expressed as
2 1/2 h2 i1/2
˙eq = ˙ij ˙ij = ˙2rr + ˙2θθ + ˙2zz
"3 3 #
2 Ė33
h A 2 Ė
33 A 2 2
i
= + 2 + − 2 + Ė33
3 2 r 2 r
2 4 A2 1/2
2 a
4 1/2
= Ė33 + = Ė 33 1 + α (11.43)
3 r4 r
where the parameter, α, is defined as
2 |A| Ω̇ 1
α := √ 2
= √ (11.44)
3 Ė33 a2 Ω 3Ė33
322 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
12 σy
− (−Ė33 )
2 3 Ė33 (1 + α2 (a/r)4 )1/2
σy
=
(1 + α2 (a/r)4 )1/2
1
= σzz − (σrr + σθθ )
2
To this end, we are in a position to link the macro-stresses, Σ11 , Σ33 − Σ11 ,
and void volume fraction, f , together in a macro yield potential.
We first link Σ11 and |Σ33 − Σ11 | with remote strain rate, Ėij .
Consider the traction boundary conditions on the surface of the void and the
surface of the RVE,
1
σrr (a) = 0, and σrr (b) = Σαα = Σ11
2
note that Σrr (b) = Σθθ (b) = 12 Σαα .
1. Integrating equilibrium equation along the radius direction yields,
Z b Z b
dσrr σθθ − σrr
Σ11 = σrr (b) − σrr (a) = dr = dr
a dr a r
Micromechanics Theory of Void Growth 323
Since,
σθθ − σrr sθθ − srr 4 σy A
= =
r r 3 Ė33 (1 + α2 (a/r)4 )1/2 r3
we have Z b
4 A dr
Σ11 = σy (11.45)
3 a Ė33 (1 + α (a/r) ) r3
2 4 1/2
2. Consider the fact that σ11 + σ22 = σrr + σθθ , and Σ11 = Σ22 = 12 Σαα ,
Z
1 1 1
Σ33 − Σ11 = Σ33 − (Σ11 + Σ22 ) = σzz − (σxx + σyy ) dV
2 V V 2
Z
1 1
= σzz − (σrr + σθθ ) dV
V V 2
Z
1 1
= szz − (srr + sθθ ) dV
V V 2
Z
1 1
= szz − (srr + sθθ ) dV
V VM 2
Recall that
1 σy
szz − (srr + sθθ ) = a 4 ,
2
(1 + α2 )1/2
r
and dV = rdrdθdz. We have
Z b
2πH σy
Σ33 − Σ11 = a 4 1/2 rdr
πb2 H a
1 + α2
r
Z b
2σy rdr
= (11.46)
b2 a
a 4 1/2
1 + α2
r
Make change of variable,
a 2
x=α : x → [α, f α], when r → [a, b] .
r
a2 Ω
where f = 2 = .
b V
Therefore,
a2 4A 2A
dx = −2α 3
dr = − √ dr ⇐ α = √
r 3Ė33 r3 3Ė33 a2
324 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
and
√
Adr 3
= − dx, (11.47)
Ė33 r3 4
r4 a2 α dx a2
rdr = − 2 dx = − , ⇐ x = α (11.48)
2a α 2 x2 r2
√
Adr 3
Reconsider (11.45) and =− dx,
Ė33 r 3 4
Z b
1 4 1 Adr
Σ11 = Σαα = σy 2 1/2
2 3 a (1 + x ) Ė33 r3
√
4 3 fα Z
dx
= − σy √
3 4 α 1 + x2
Thereby, Z α
Σαα σy dx
=√ √ (11.49)
2 3 f α 1 + x2
We then find that the in-plane hydrostatic stress can be written as
√
3 Σαα h α + √1 + α 2 i
= log p
2 σy f α + 1 + (f α)2 (11.50)
αa2 dx
Reconsider Eq. (11.46) and rdr = − ,
2 x2
2σy b
Z
rdr
Σ33 − Σ11 = 2 2 1/2
b a (1 + x )
2σ αa2 Z f α dx
y
= 2
− √
b 2 α x 1 + x2
2
Z α
dx
= f ασy √
2 1 + x2
fα x
We can then link the deriatoric macro-stress with macro-strain rate and void
volume fraction,
Σeq p p
= 1 + f 2 α2 − f 1 + α2
σy (11.51)
Micromechanics Theory of Void Growth 325
Denote that
√
3 Σαα
A1 =
2 σy
Σeq
A2 =
σy
p
A3 = α + 1 + α2
p
A4 = f α + 1 + f 2 α2
A2 exp(A1 )
A3 =
1 − f exp(A1 ) (11.56)
A2
A4 =
1 − f exp(A1 ) (11.57)
326 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
A23 − 1 A2 − 1
2α = = 4
A3 f A4 (11.61)
Substituting
A2 exp(A1 )
A3 =
1 − f exp(A1 )
A2
A4 =
1 − f exp(A1 )
into (11.61), we obtain the following identity,
A23 − 1 A2 − 1 A2 exp(2A1 ) − (1 − f exp(A1 ))2 A2 exp(A1 )
= 4 ⇒ 2 =
A3 f A4 A22 − (1 − f exp(A1 ))2 f A2
Rewrite the above equation,
f A22 exp(2A1 ) − f (1 − f exp(A1 ))2
= A22 exp(A1 ) − exp(A1 )(1 − f exp(A1 ))2
⇒ A22 exp(A1 )(1 − f exp(A1 )) = (1 − f exp(A1 ))2 (exp(A1 ) − f )
which leads to
A22 = (1 − f exp(A1 ))(1 − f exp(A1 ))
h i
= 1 + f 2 − f exp(A1 ) + exp(−A1 )
= 1 + f 2 − 2f cosh A1
We finally link A1 and A2 in a single equation.
Micromechanics Theory of Void Growth 327
Σ2eq √3 Σ
αα
F (Σeq , Σαα , f ) = 2 + 2f cosh − (1 + f 2 ) = 0 .
σy 2 σy (11.62)
Therefore,
√3 Σ
αα
p p
sinh = α( 1 + f 2 α2 − f 1 + α2 ) (11.63)
2 σy
Consider
Ω̇ 1
α = √ (11.64)
Ω 3Ė33
Σeq p p
= 1 + f 2 α2 − f 1 + α2 (11.65)
σy
Eq. (11.63) can be rewritten as
√3 Σ Ω̇ 1 Σeq
αα
sinh = √
2 σy Ω 3Ė33 σy
or
Ω̇ √ σ √3 Σ
y αα
= 3Ė33 sinh
Ω Σeq 2 σy
Ω̇
Considering the fact f˙ = f (1 − f ), we recover the McClintock solution,
Ω
√ σ √3 Σ
y αα
f˙ = 3f (1 − f )Ė33 sinh
Σeq 2 σy (11.66)
328 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
11.4 Exercise
References
Ablowitz, M. and Fokas, A. S. (1997), Complex Variables: Introduction and Applications, Cam-
bridge University Press.
Cottrell, A. H. (1953), Dislocations and Plastic Flow in Crystals, Oxford University Press.
Eshelby, J. D. (1957) “The determination of the elastic field of an ellipsoidal inclusion and
related problems,” Proceedings of Royal Society of London, A 241, pp. 376-396.
Eshelby, J. D. (1959), “The elastic field outside an ellipsoidal inclusion,” Proceedings of Royal
Society of London, A 252, pp. 561-569.
Hashin, Z. and Shtrikman, S. [1961], “Note on a variational approach to the theory of composite
elastic materials,” The Franklin Institute Laboratories, 271, pp. 336-341.
Hashin, Z. and Shtrikman, S. [1962], “On some variational principlrd in anisotropic and nonho-
mogeneous elasticity,” Journal of Mechanics and Physics of Solids, 10, pp. 335-342.
Hashin, Z. and Shtrikman, S. [1962], “The elastic moduli of heterogeneous materials,” Journal
of Applied Mechanics, pp. 143-148.
Hashin, Z. (1965), “On elastic behaviour of fibre reinforced materials of arbitrary transverse
phase geometry,” Journal of Mechanics and Physics of Solids, 13, pp. 119-134.
Hirth, J. P. and Lothe, J. (1982). Theory of Dislocations, Krieger
Love, A. E. H. (1926). A Treatise on the Mathematical Theory of Elasticity, Fourth Edition,
Dover Publication, New York.
Malvern, L. E. (1969). Introduction to the Mechanics of a Continuous Mdedium, Prentice-Hall.
Mura, T. (1987). Micromechanics of Defects in Solids, Second, Revised Edition, Kluwer Aca-
demic Pub. Dordrecht/Boston/London
Nemat-Nasser, S. and Hori, M. (1999), Micromechanics: overall properties of heterogeneous
materials, Elsevier, Amsterdam-Lausanne-New York
Sneddon, I. N. (1950), Fourier Transforms, Dover Publication, Inc. New York.
Sokolnikoff, I. S. (1956). Mathematical Theory of Elasticity, 2n edition, McGraw-Hill
330 INTRODUCTION TO MICROMECHANICS AND NANOMECHANICS
Talbot, D.R.S. and Willis, J.R. [1985], “Variational principles for inhomogeneous non-linear
media,” IMA Journal of Applied Mathematics, 35, pp. 39-54.
Talbot, D.R.S. and Willis, J.R. [1998], “Upper and lower bounds for the overall response of an
elastoplastic composite,” Mechanics of Materials, 28, pp. 1-8.
Timoshenko, S. and Goodier, J. N. (1951). Theory of Elasticity, 2nd ed. McGraw-Hill
Titchmarsh, E. C. (1948). Introduction to the Theory of Fourier Integrals, Oxford.
REFERENCES 331