Muthukumar T - Homogenization
Muthukumar T - Homogenization
attains its minimum, where u(x) is the solution of the Dirichlet problem
−div(A(x)∇u(x)) = f (x) in Ω
(0.0.1)
u =0 on ∂Ω,
A(x) = a(x)I and (
α on Ω1
a(x) =
β on Ω \ Ω1 .
i
CHAPTER 0. PREFACE ii
Preface i
Notations vii
1 Asymptotic Expansion 1
1.1 Periodically Oscillating Functions . . . . . . . . . . . . . . . . 1
1.2 Second Order Elliptic Equation . . . . . . . . . . . . . . . . . 3
1.3 Periodic Boundary Conditions . . . . . . . . . . . . . . . . . . 4
1.4 Periodic Composite Material . . . . . . . . . . . . . . . . . . . 5
1.5 Asymptotic Expansion in Two Scales . . . . . . . . . . . . . . 7
2 Two-Scale Convergence 19
2.1 Vector-valued Function Spaces . . . . . . . . . . . . . . . . . . 19
2.2 Two-scale Convergence . . . . . . . . . . . . . . . . . . . . . . 24
2.3 Classical Definition of Two-Scale Convergence . . . . . . . . . 32
2.4 Homogenization of Second Order Linear Elliptic Problems . . 33
2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3 H-Convergence 39
3.1 Coercive Operators . . . . . . . . . . . . . . . . . . . . . . . . 39
3.2 H-Convergence . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.3 Correctors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
3.4 Generalised Energy Convergence . . . . . . . . . . . . . . . . . 58
3.5 Optimal Bounds . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4 Γ-Convergence 63
4.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.2 Direct Method of Calculus of Variation . . . . . . . . . . . . . 63
v
CONTENTS vi
5 Bloch-Floquet Homogenization 77
5.1 Fourier Transform . . . . . . . . . . . . . . . . . . . . . . . . . 77
5.2 Schrödinger Operator with Periodic Potential . . . . . . . . . 78
5.2.1 Direct Integral Decomposition . . . . . . . . . . . . . . 79
5.3 Bloch Periodic Functions . . . . . . . . . . . . . . . . . . . . . 83
5.4 Bloch Transform . . . . . . . . . . . . . . . . . . . . . . . . . 85
5.4.1 Spectrum of Elliptic Operator . . . . . . . . . . . . . . 88
5.4.2 Regularity of λm (η) and φ1 (·, η) . . . . . . . . . . . . . 90
5.4.3 Taylor Expansion of Ground State . . . . . . . . . . . 92
5.5 Homogenization of Second order Elliptic Operator . . . . . . . 94
Appendices 101
Bibliography 103
Index 109
Notations
Symbols
M (α, β, Ω) denotes, for 0 < α < β, the class of all n × n matrices, A = A(x),
with L∞ (Ω) entries such that,
t
A denotes the transpose of a matrix A
Function Spaces
D(Ω) or Cc∞ (Ω) is the class of all infinitely differentiable functions on Ω with
compact support
H01 (Ω) is the closure of D(Ω) in W 1,2 (Ω) = (H 1 (Ω)) and its norm is denoted
by k.kH01 (Ω)
H −1 (Ω) is the dual space of H01 (Ω) and its norm is denoted by k · kH −1 (Ω)
vii
NOTATIONS viii
L∞ (Ω) is the space of all essentially bounded measurable functions and its
norm is denoted by k.k∞,Ω
Lp (Ω) is the space of all p-summable measurable functions and its norm is
denoted by k.kp,Ω (1 ≤ p < ∞)
Lp (Ω; X)
R denotes the class of all measurable functions f : Ω → X such that
p
Ω
kf (x)kX < ∞, where X is a Banach space
Lpper (Y ) denotes the class of Y -periodic functions in Lploc (Rn ) and its norm
denoted by k.kp,Y (1 ≤ p ≤ ∞)
W m,p (Ω) is the collection of all Lp (Ω) functions such that all distributional
derivatives upto order m are also in Lp (Ω) and its norm is denoted by
k.km,p,Ω
m,p
Wper (Y ) denotes the class of Y -periodic functions in W m,p (Rn ) and its norm
denoted by k · km,p,Y
General Conventions
Asymptotic Expansion
For simplicity, we shall, henceforth, take the reference cell to be the unit
cube of Rn , i.e., Y = [0, 1]n . This is only for simplicity and to avoid carrying
the measure of Y , |Y |, in our calculations. Note that Rn = ∪k∈Zn (k + Y ),
is a disjoint union. Any function f : Y → R, defined a.e. on Y , may be
extended a.e. to Rn as a Y -periodic function. Let Lpper (Y ) denote the set
of all Lploc (Rn ) which are Y -periodic equipped with the norm of Lp (Y ). For
any f ∈ Lpper (Y ) and ε > 0, we may define a new function fε : Rn → R as
1
CHAPTER 1. ASYMPTOTIC EXPANSION 2
extended to all of R. Define, for any 0 < ε < 1, the new function fε (x) =
f (x/ε) on R. Note that the number of points of jump discontinuity for f in
[0, 1] is only one at y = 1/2. However, fε has more than one point of jump
discontinuity in [0, 1]. For instance, for ε = 1/2, the function fε has three
points of jump discontinuity, at x = 1/4, 1/2, 3/4, in [0, 1].
Example 1.2. Consider f (y) = sin(2πy) on [0, 1] extended to all of R. Define
fε (x) = sin(2πx/ε) on R. For any 0 < ε < 1, we see that the number of
oscillations on [0, 1] is increased for fε . For instance, for ε = 1/2, fε has twice
the number of oscillations, as that of f , on [0, 1].
Theorem 1.1.2. Let 1 ≤ p ≤ +∞ and f ∈ Lpper (Y ). Then fε (x) = f (x/ε),
for 0 < ε < 1, is bounded in any open cell R that contains any translation of
Y , i.e.,
|R|
kfε kpp,R ≤ C0 kf kpp,Y
|Y |
where C0 depends only on n, the dimension of Euclidean space.
Theorem 1.1.3. Let 1 ≤ p ≤ +∞, f ∈ Lpper (Y ) and set fε (x) = f (x/ε), for
0 < ε < 1, on Rn . For 1 ≤ p < ∞,
Z
1
fε * f (y) dy weakly in Lp (Ω)
|Y | Y
for any bounded open subset Ω ⊂ Rn . If p = ∞, then
Z
1
fε * f (y) dy weak-* in L∞ (Ω).
|Y | Y
1
constant functions have no oscillations
CHAPTER 1. ASYMPTOTIC EXPANSION 3
i.e.,
J(u) = min 1
J(v).
v∈H0 (Ω)
CHAPTER 1. ASYMPTOTIC EXPANSION 4
Theorem 1.3.1. Let Y be unit open cell and let aij ∈ L∞ (Ω) such that the
matrix A(y) = (aij (y)) is elliptic with ellipticity constant α > 0. For any
f ∈ (Wper (Y ))? , there is a unique weak solution u ∈ Wper (Y ) satisfying
Z
A∇u · ∇v dx = hf, vi(Wper (Y ))? ,Wper (Y ) ∀v ∈ Wper (Y ).
Y
CHAPTER 1. ASYMPTOTIC EXPANSION 5
εY
y
x
Figure 1.1: partition of Ω into ε-cells
The condition given above is called the ellipticity condition. We extend aij
to all of Rn , and for each 0 < ε < 1, we define the function aεij : Rn → R as
x
aεij (x) = aij a.e. x ∈ Rn
ε
and the n × n matrix Aε (x) = (aεij (x)) is in M (α, β, Ω). Thus, the Dirichlet
problem for the composite material Ω is given as, for a given f ∈ H −1 (Ω),
−div(Aε (x)∇uε (x)) = f (x) in Ω
(1.4.1)
uε = 0 on ∂Ω.
CHAPTER 1. ASYMPTOTIC EXPANSION 7
By Lax-Milgram result, there exists a unique solution uε ∈ H01 (Ω) such that
Z
Aε (x)∇uε (x) · ∇v(x) dx = hf, viH −1 (Ω),H 1 (Ω) , ∀v ∈ H01 (Ω)
0
Ω
and kuε kH01 (Ω) ≤ 1/αkf kH −1 (Ω) . Computing the solution, numerically, is
stable if the size of the grid is chosen smaller than ε. But for composite
materials, ε is very small and choosing grid smaller than ε leads to impossible
computable situation. Therefore, we study the limiting case, as ε → 0, of
the Dirichlet problem (1.4.1).
Note that in the periodic set-up any x ∈ Ω has two reprsentations. One is
the macroscale representation x and the other is that x is in some translation
of the εY cell having the form x = εy for some y ∈ Y . Thus, any x ∈ Ω may
take two representations each in Ω and Y as x and y = xε , respectively. Thus,
the behaviour of a periodic composite material Ω, as given in (1.4.1), involves
two scales, viz., the “macroscopic or slow” scale x ∈ Ω and the “microscopic
or fast” scale y = x/ε ∈ Y . We intend to find uε (x) that solves (1.4.1).
Thus, our model suggests that uε depends on both the slow variable x and
fast variable y = x/ε, viewed as independent variables. This suggests us to
seek uε (x), with x ∈ Ω, in the form
x x x
uε (x) = u0 x, + εu1 x, + ε2 u2 x, + ..., (1.5.1)
ε ε ε
where ui (x, y) are functions which are Y -periodic in the y-variable. The
partial derivative of a function φε (x) := φ(x, x/ε) is given as
A2 u0 = 0 (1.5.2)
A2 u1 + A1 u0 = 0 (1.5.3)
A2 u2 + A1 u1 + A0 u0 = f (x) (1.5.4)
A2 um+2 + A1 um+1 + A0 um = 0 ∀m ≥ 1. (1.5.5)
Consider
Z
A1 u(x) dy = hA1 u(x), 1i
Y
Z X n
∂ ∂u(x)
= − aij (y) · 1 dy
Y i,j=1 ∂yi ∂xj
n Z
X ∂u(x) ∂1
= aij (y) = 0.
i,j=1
∂xj Y ∂yi
or equivalently,
R j − yj ) = 0 in Y
div(A(y)∇(χ
1
χ (y) dy = 0
|Y | Y j
χj − yj is Y -periodic.
CHAPTER 1. ASYMPTOTIC EXPANSION 10
n
X ∂u(x)
A2 u1 (x, y) + A1 u(x) = A2 u1 (x, y) + A2 χj (y)
j=1
∂xj
n
!
X ∂u(x)
= A2 u1 (x, y) + A2 χj (y)
j=1
∂xj
n
!
X ∂u(x)
= A2 u1 (x, y) + χj (y) .
j=1
∂xj
Therefore, we have
n
X ∂u(x)
u1 (x, y) = − χj (y) + ũ(x)
j=1
∂xj
A2 u2 (x, y) = f (x) − A1 u1 (x, y) − A0 u(x) in Y
u2 (·, y) is Y -periodic in y
R
u (x, y) dy = 0.
Y 2
Z Z
f (x) = (A1 u1 (x, y) + A0 u(x)) dy.
Y Y
CHAPTER 1. ASYMPTOTIC EXPANSION 11
n
(Z " n
!# )
X ∂ ∂ X ∂u(x)
= − aij (y) ũ(x) − χk (y) dy
i,j=1
∂xi Y ∂yj k=1
∂xk
n Z
X ∂ ∂χk (y) ∂u(x)
= aij (y) dy
i,j,k=1
∂xi Y ∂yj ∂xk
n Z 2
X ∂χk (y) ∂ u(x)
= aij (y) dy .
i,j,k=1 Y ∂yj ∂xi ∂xk
Next, we consider
Z Z X n
∂ ∂u(x)
A0 u(x) dy = − aij (y) dy
Y Y i,j=1 ∂xi ∂xj
n Z 2
X ∂ u(x)
= − aij (y) dy .
i,j=1 Y ∂x i ∂x j
For convenience, we rewrite the above relation by replacing the index j with
k to get
Z n Z 2
X ∂ u(x)
A0 u(x) dy = − aik (y) dy .
Y i,k=1 Y ∂xi ∂xk
where !
Z n
1 X ∂χk (y)
a0ik = aik (y) − aij (y) dy. (1.5.7)
|Y | Y j=1
∂yj
α|ξ|2
Z
≥ |∇y η|2 dy = α0 |ξ|2
|Y | Y
where Z
α
α0 = |∇y η|2 dy
|Y | Y
Pn ξi
and η(y) := i=1 |ξ| (yi − χi (y)). To show the ellipticity of (a0ij ), it is enough
to show that α0 > 0. On the contrary, suppose α0 = 0 then |∇y η(y)| = 0
for a.e. y ∈ Y . Since Y is connected, η is constant on Y , say C0 . Thus,
CHAPTER 1. ASYMPTOTIC EXPANSION 13
Pn Pn
i=1 ξi yi = i=1 ξi χi (y) + |ξ|C0 . Since χi is such that its average is zero, we
obtain
n Z
X 1
yi dy = |ξ|C0 .
i=1
|Y | Y
From the result proved abvoe and the fact that f ∈ H −1 (Ω), the equation
n
X ∂ 2 u(x)
− a0ik = f (x) (1.5.8)
i,k=1
∂xi ∂xk
n
" n #
X X ∂χk (y) ∂ 2 u(x)
A1 u1 (x, y) + A0 u(x) = aij (y) − aik (y)
i,k=1 j=1
∂yj ∂xi ∂xk
n
X ∂ ∂u1
− aij (y)
i,j=1
∂yi ∂xj
CHAPTER 1. ASYMPTOTIC EXPANSION 14
Now, using the homogenized equation (1.5.8) for f in the above relation, we
get A2 u2 (x, y) =
n
" n #
∂ 2 u(x)
X X ∂χk (y) ∂[aji (y)χk (y)]
− aij (y) + − aik (y) + a0ik
i,k=1 j=1
∂yj ∂yj ∂xi ∂xk
n
X ∂aij (y) ∂ ũ(x)
+
i,j=1
∂yi ∂xj
Therefore, we have
n n
X ∂ 2 u(x) X ∂ ũ(x)
u2 (x, y) = θik − χj (y) + ũ2 (x)
i,k=1
∂xi ∂xk j=1 ∂xj
for some function ũ2 (x). This way one can proceed for all values of uk , using
the recurrence relation (1.5.5). Finally, substituting the values of uk (x, y) in
(1.5.1), we have
n
!
X ∂u(x)
uε (x) = u(x) + ε − χj (y) + ũ(x)
j=1
∂x j
n n
!
2
X ∂ u(x) X ∂ ũ(x)
+ε2 θik − χj (y) + ũ2 (x) + . . .
i,k=1
∂xi ∂xk j=1 ∂xj
uε (x) → u(x) as ε → 0
Example 1.3. Let us understand the above method in the one dimensional
case. Let a : [0, 1] → R be a function such that 0 < α ≤ a(x) < β a.e.
in [0, 1]. Thus, a satisfies the ellipticity condition and is in L∞ (Y ), and is
extended periodically to all of R and aε (x) = a(x/ε), for x ∈ (a, b) ⊂ R. Let
Y = (0, 1) and the equation (1.4.1) takes the form
and, therefore, a(y)χ0 (y) = a(y) + c, for some constant c. This first order
differential equation will admit a periodic solution iff
Z Z
c 1
1+ dy = 1 + c dy = 0.
Y a(y) Y a(y)
Note that due to the ellipticity condition, a(y) > 0 for all y, and hence there
is no division by zero. The effective coefficient is given by formula (1.5.7)
Z Z
dχ(y)
a0 = a(y) − a(y) dy = (a(y) − a(y) − c) dy = −c.
Y dy Y
CHAPTER 1. ASYMPTOTIC EXPANSION 17
Hence, Z −1
1
a0 = dy .
Y a(y)
The interesting fact to be observed here is that the effective coefficient a0 is
the inverse of the weak limit in Lp (Ω) of 1/aε rather that the weak limit of
aε , as one would expect. So, the homogenized coefficient is not always same
as taking the averages.
We caution that for the homogenized equation (1.5.8) to be solvable we
need to check that the effective coefficients a0ik satisfy the ellipticity condition
and are bounded. We shall address these questions after we give a precise
mathematical formulation of deriving the homogenized equation.
CHAPTER 1. ASYMPTOTIC EXPANSION 18
Chapter 2
Two-Scale Convergence
The aim of this chapter is to give a rigorous treatment of the formal asymp-
totic expansion of homogenization problems with periodic coefficients, dis-
cussed in previous chapters. The two-scale method justifies the formal expan-
sion. The notion of two-scale convergence was introduced by G. Nguetseng
(cf. [Ngu89]) in 1989 and further developed by G. Allaire (cf. [All92, All94,
LNW02]).
In Chapter 1, we studied (1.4.1) by introducing a formal two-scale asymp-
totic expansion for the solution uε , i.e.,
∞
X x
i
uε (x) = ε ui x, .
i=0
ε
The above power series expansion was proposed, expecting that the solution
uε will exhibit a two-scale oscillation, viz., in x and y = xε variables. This is
due to the presence of two scales in (1.4.1). The formal two-scale asymptotic
expansion method is not rigorous. Two-scale convergence method incorpo-
rates the lessons learnt from the two-scale asymptotic expansion and gives
rise to a rigorous justification of the homogenization process. In this chap-
ter, we shall consider more general coefficients A(x, y) = (aij (x, y)) defined
on Ω × Y instead of A(y).
19
CHAPTER 2. TWO-SCALE CONVERGENCE 20
kf (x)kpX < ∞.
R
class of all measurable functions f : Ω → X such that Ω
Proof. Suppose f ∈ L1 (Ω; Cper (Y )), then (a) and (c) are obvious from def-
initions and it only remains to prove (b). Since the map f : Ω → Cper (Y )
is measurable, by Pettis’ theorem, x 7→ hG, f (x)i is measurable for every
G ∈ [Cper (Y )]? . In particular, for
R any fixed y ∈ Y , choose G to be the Dirac
measure δy at y, i.e., hδy , gi = Y g(t) dδy = g(y), for all g ∈ Cper (Y ). Thus,
is measurable.
Conversely, if (a), (b) and (c) are satisfied then it only remains to prove
that the map f : Ω → Cper (Y ) is measurable. By Pettis’ theorem, it is
CHAPTER 2. TWO-SCALE CONVERGENCE 21
enough to prove that x 7→ hG, f (x)i is measurable, for every G ∈ [Cper (Y )]? .
Without loss of generality, assume G to be a positive functional because any
G can be split into positive and negative functionals. By RieszR representation,
there is a unique positive measure µG such that hG, gi = Y g dµG , for all
g ∈ Cper (Y ). We now approximate G by a sequence of functionals Gk which
are finite linear combination of Dirac measures. Let {Yi }k1 be a partition of
Y into k-disjoint cubes of side length k1 and λi := µG (Yi ). Let yi ∈ Yi , χYi
be the characteristic function of Yi in Y , extended periodically to Rn and
δi := δyi be the Dirac measure at yi . Define Gk as,
k Z
X k
X
hGk , gi := g(y) d(λi δi ) = λi g(yi ).
i=1 Y i=1
k
X k
X
hGk , f (x)i = λi f (x, yi ) = f (x, yi )µG (Yi )
i=1 i=1
Z k
!
X
= f (x, yi )χYi (y) dµG (y).
Y i=1
For eachP x ∈ Ω, let Ak (x) be the class of all simple functions of the form
sx (y) = ki=1 αi (x)χYi (y) and sx (y) ≤ f (x, y), for all y ∈ Y . Also, let S(x)
be the class of all simple functions satisfying sx (y) ≤ f (x, y), for all y ∈ Y .
CHAPTER 2. TWO-SCALE CONVERGENCE 22
Then
Z k
!
X
lim hGk , f (x)i = lim f (x, yi )χYi (y) dµG (y)
k→∞ k→∞ Y i=1
" Z #
≤ lim sup sx (y) dµG (y)
k→∞ sx ∈Ak (x) Y
Z
≤ sup sx (y) dµG (y)
sx ∈S(x) Y
Z
≤ f (x, y) dµG (y)
Y
= hG, f (x)i.
By taking −f instead of f in the above argument and by linearity of duality,
we obtain the reverse inequality and, hence, the equality.
For 1 ≤ p < ∞, let Xp (Ω; Y ) generically denote one of the follow-
ing spaces: Lp (Ω; Cper (Y )), Lpper (Y ; C(Ω)), C(Ω̄; Cper (Y )). The demand of
smoothness in one of the variables is mandatory. The space Xp (Ω; Y ) is a
separable Banach space and Xp (Ω; Y ) is dense in Lp (Ω × Y ).
Theorem 2.1.4. Let 1 ≤ p < ∞. For any φ ∈ Lp (Ω; Cper (Y )), the functions
φε (x) := φ x, xε are measurable in x and satisfies:
(a)
kφε kp,Ω ≤ kφkLp (Ω;Cper (Y )) (2.1.1)
and, hence, φε ∈ Lp (Ω);
(b) Z
ε 1
φ * φ(·, y) dy weakly in Lp (Ω). (2.1.2)
|Y | Y
1
In particular, for p = 2, kφε k22,Ω → |Y |
kφk22,Ω×Y .
Proof. We shall prove the result only for Xp (Ω; Y ) = L1 (Ω; Cper (Y )) because
the proof is similar in all other cases. By Theorem 2.1.3, the functions φε
are Caratheodory (cf. [All92, ET74]) and hence they are measurable. The
inequality (2.1.1) is easy to conclude because
Z x Z
ε
kφ k1,Ω = |φ x, | dx ≤ sup |φ(x, y)| dx = kφkL1 (Ω;Cper (Y )) .
Ω ε Ω y∈Y
CHAPTER 2. TWO-SCALE CONVERGENCE 23
We shall now prove (2.1.2). Consider the partition {Yi }k1 of Y and yi ∈ Yi as
in Theorem 2.1.3. Let χYi be the characteristic function of Yi in Y , extended
periodically to Rn . Define the step functions
k
X
φk (x, y) = φ(x, yi )χYi (y).
i=1
But x Z
1
χYi * χYi (y) dy weak-* in L∞ (Ω).
ε |Y | Y
Thus, (2.1.2) is true for step functions. But, for each fixed k ∈ N and
ψ ∈ L∞ (Ω),
Z Z Z
ε 1
φ (x) − φ(x, y) dy ψ(x) dx ≤ |φε (x) − φεk (x)| |ψ(x)| dx
Ω |Y | Y
ZΩ Z
ε 1
+ φk − φk dy |ψ| dx
Ω |Y | Y
Z
1
+ |φk − φ| |ψ(x)| dy dx.
|Y | Ω×Y
Thus,
Z Z Z
ε 1
lim φ − φ dy ψ dx ≤ 2 sup |φ(x, y) − φk (x, y)| |ψ(x)| dx
ε→0 Ω |Y | Y Ω y∈Y
= 2 k(φ − φk )ψkL1 (Ω;Cper (Y )) .
Define
gk (x) := sup |φ(x, y) − φk (x, y)| |ψ(x)|.
y∈Y
Theorem 2.1.5 (cf. [BM]). Let 1 ≤ p < ∞. Suppose φ(x, y) = φ1 (x)φ2 (y)
such that φ1 ∈ Ls (Ω) and φ2 ∈ Ltper (Y ) with 1 ≤ s, t < ∞ and 1s + 1t = p1 .
Then φε (x) = φ x, xε ∈ Lp (Ω) and
Z
ε φ1 (·)
φ * φ2 (y) dy weakly in Lp (Ω).
|Y | Y
periodic in the second variable. For any φ ∈ Lq (Ω; Cper (Y )), set ψ(x, y, z) :=
u(x, z)φ(x, y). Recall that if ψ : Ω × Rn × Rn → R is a function which is
Y -periodic in both second variable and third variable, then
x x Z Z
1
ψ x, , 2 * 2
ψ(x, y, z) dz dy weakly in Lp (Ω).
ε ε |Y | Y Y
Therefore,
Z Z Z Z
x x 1
u x, 2 φ x, dx → u(x, z)φ(x, y) dz dy dx.
Ω ε ε |Y |2 Ω Y Y
Equivalently, Z
x 2s 1
u x, 2 * u(x, z) dz.
ε |Y | Y
Thus, the two-scale limit is same as the Lp weak limit. More generally, if
Yi = Y , for all i = 1, 2, . . . , k,
x Z Z
x 2s 1
u x, 2 , . . . , k * ... u(x, y2 , . . . , yk ) dy2 . . . dyk
ε ε |Y |k−1 Y2 Yk
Remark 2.2.3. Recall that uε * u weakly in Lp (Ω) if, for all φ ∈ Lq (Ω),
Z Z
uε (x)φ(x) dx → u(x)φ(x) dx.
Ω Ω
Thus, the usual weak convergence in Lp (Ω) hides (averages out) the effect
of oscillations in uε . In order to capture the oscillations of the form xε ,
one has to treat uε with test functions of the form φ x, xε . This was the
motivation behind the definition of two-scale convergence. Also, note that,
as seen in the above example, the test function φ x, xε is not good enough
The basic idea is that one has to treat with test functions with same order
of oscillations. This is called the multi-scale or reiterated homogenization.
Remark 2.2.4. Suppose uε admits an asymptotic expansion
∞
X x
uε (x) = εi ui x,
i=0
ε
Z
uε (x)φε (x) − u(x)φ̄(x) dx
≤ kuε − ukp,Ω kφε kq,Ω
Ω
Z
u(x) φε (x) − φ̄(x) dx .
+
Ω
L2 ([0, 1] × [0, 1]), defined as u(x, y) = sin(2πy), and define the sequence
uε (x) := sin 2πxε
in L2 [0, 1]. Because {uε } converge weakly to 0 in L2 [0, 1]
(periodic oscillating function weakly converges to average), if it strong con-
verges then the limit must be 0. But k sin(2πx/ε)k2,[0,1] = 1/2 and, hence, do
not strongly converge. The two-scale limit of the sequence uε (x) := sin 2πx ε
is u(x, y) = sin(2πy) on [0, 1] × [0, 1].
p 2s p
Theorem 2.2.6.
1
R For any sequence uε ⊂pL (Ω), if uε * u with u ∈ L (Ω×Y )
then uε * |Y | Y u(x, y)dy weakly in L (Ω). In particular, if the two-scale
limit u is independent of y then the two-scale limit and weak limit coincide.
2s
Proof. Let uε * u. Then, in particular, for any φ ∈ Lq (Ω) ⊂ Lq (Ω; Cper (Y ))
(φ independent of y),
Z Z Z
1
uε (x)φ(x) dx → u(x, y)φ(x) dy dx.
Ω |Y | Ω Y
Thus, Z
1
uε * u(·, y)dy weakly in Lp (Ω).
|Y | Y
If u(x, y) = u(x) then |Y1 | Y u(x) dy = u(x). Thus, weak limit and two-scale
R
This sequence converges weakly to zero in L2 ([0, 1] but does not two-scale
converge.
The last inequality follows from Theorem 2.1.4. Since {uε } is bounded
in Lp (Ω), there is a constant C0 > 0 (independent of ε) such that
kuε kp,Ω ≤ C0 . Thus, the sequence {Lε } is bounded in the dual of
Lq (Ω; Cper (Y )).
for all φ such that φ(x, y) = φ1 (x)φ2 (y) where φ1 ∈ Ls (Ω) and φ2 ∈ Lt (Ω)
with 1 ≤ s, t < ∞ and 1s + 1t = 1q .
This implies the first inequality of (2.2.2). The second inequality in (2.2.2)
follows from Jensen’s inequality,
Z Z p Z Z
p
p
|Y | kūkp,Ω = u(x, y) dy dx ≤ |u(x, y)|p dy dx = kukpp,Ω×Y .
Ω Y Ω Y
m,p
Let Wper (Y ) denote the class of functions in W m,p (Rn ) which are Y -
periodic.
Theorem 2.2.12 (Compactness in W 1,p ). For any given 1 ≤ p < ∞, let uε
be a bounded sequence in W 1,p (Ω). Then there is exists a u ∈ W 1,p (Ω) and
u1 ∈ Lp (Ω; Wper
1,p
(Y )) such that, for a subsequence (still denoted by ε),
∞
for all Φ ∈ [D(Ω; Cper (Y ))]n such that divy Φ = 0. By a classical result (cf.
[Tem79, GR81]), v(x, y) − ∇u(x) is a gradient with respect to y, i.e., there
is a u1 ∈ Lp (Ω; Wper
1,p
(Y )) such that
∞
Example 2.6. The choice of C(Ω; Cper (Y )) as test function space will, also,
not yield any better result. Let ũε be the periodic extension of uε , defined
above, to all of R. Define vε : (0, 1) → R as
(
ũε ( x ) if 41 < x < 34
vε (x) =
0 otherwise.
∞
Then, for all φ ∈ C((0, 1); Cper (0, 1)),
Z 1 x Z 1Z 1
vε (x)φ x, dx → u(x, y)φ(x, y) dy dx
0 ε 0 0
where (
1 if 41 < x < 3
4
u(x, y) =
0 otherwise.
Note that {vε } is not bounded in L2 (0, 1).
Theorem 2.3.1. Let {uε } be a bounded sequence in Lp (Ω) and
Z x Z Z
1 ∞
uε (x)φ x, dx → u(x, y)φ(x, y) dy dx ∀φ ∈ D(Ω; Cper (Y )).
Ω ε |Y | Ω Y
2s
Then uε * u.
Proof. The idea is to approximate φ ∈ Lq (Ω; Cper (Y )) by a sequence φk ∈
∞
D(Ω; Cper (Y )) and use the uniform bound of {uε }.
Theorem 2.3.2 (Compactness). Let {uε } be a bounded sequence in Lp (Ω).
Then, along a subsequence,
Z x Z Z
uε (x)φ x, dx → u(x, y)φ(x, y) dy dx ∀φ ∈ Lqper (Y ; C(Ω̄)).
Ω ε Ω Y
∞
Proof. Use the density of D(Ω; Cper (Y )) in Lqper (Y ; C(Ω̄)).
know there exists a unique solution uε ∈ H01 (Ω), by Lax-Milgram result, such
that
Z
Aε (x)∇uε (x) · ∇v(x) dx = hf, viH −1 (Ω),H 1 (Ω) , ∀v ∈ H01 (Ω)
0
Ω
and kuε kH01 (Ω) ≤ 1/αkf kH −1 (Ω) . Since uε is uniformly bounded in H01 (Ω),
by Theorem 2.2.12, there exists a u ∈ H01 (Ω) and u1 ∈ L2 (Ω; Hper1
(Y )) such
that, for a subsequence
uε * u
weakly in H01 (Ω) and strongly in L2 (Ω);
2s
uε * u in L2 (Ω); (2.4.2)
∇u * 2s 2 n
ε ∇u + ∇y u1 in [L (Ω)] .
∞
Consider the test functions φ ∈ D(Ω) and φ1 ∈ D(Ω; Cper (Y )) and choose
x
v(x) = φ(x) + εφ1 x, ε in the weak formulation above to obtain
Z x D x E
A x, ∇uε · (∇φ + ε∇φ1 + ∇y φ1 ) dx = f, φ + εφ1 x, .
Ω ε ε H −1 (Ω),H01 (Ω)
Note that φ + εφε1 * φ weakly in H01 (Ω). Thus, the term in RHS converges,
i.e.,
hf, φ + εφε1 iH −1 (Ω),H 1 (Ω) → hf, φiH −1 (Ω),H01 (Ω) .
0
Note that {Aε (x)∇uε (x)} is bounded in [L2 (Ω)]n and, by (2.1.1), {∇φε1 } is
bounded in [L2 (Ω)]n . Using Hölder’s inequality, we obtain
Z x
A x, ∇uε · ε∇φ1 dx ≤ Cε.
Ω ε
Therefore, Z x
lim A x, ∇uε · ε∇φ1 dx = 0.
ε→0 Ω ε
Let us denote ψ(x, y) := t A(x, y)(∇φ(x) + ∇y φ1 (x, y)). Then
Z x x Z x
A x, ∇uε · ∇φ(x) + ∇y φ1 x, dx = ∇uε · ψ x, dx.
Ω ε ε Ω ε
Since ψ is a two-scale test function, one may pass to limit in the RHS above.
Thus,
Z Z
ε ε→0 1
Aε ∇uε · (∇φ + ∇y φ1 ) dx → [∇u + ∇y u1 (x, y)] · ψ(x, y) dy dx
Ω |Y | Ω×Y
CHAPTER 2. TWO-SCALE CONVERGENCE 35
and, therefore,
Z
1
A(x, y)(∇u + ∇y u1 ) · (∇φ + ∇y φ1 (x, y)) dy dx = hf, φiH −1 (Ω),H01 (Ω)
|Y | Ω×Y
(2.4.3)
∞
for all φ ∈ D(Ω) and φ1 ∈ D(Ω; Cper (Y )). Thus, by density, for all φ ∈ H01 (Ω)
and φ1 ∈ L2 (Ω; Hper
1
(Y )). In particular, by choosing φ1 ≡ 0, we get
( h i
1
R
−divx |Y | Y
A(x, y)(∇u(x) + ∇y u1 (x, y)) dy = f in Ω
(2.4.4)
u =0 on Ω
Both (2.4.4) and (2.4.5) are called the coupled two-scale homogenized system
of equations. The system (2.4.4) and (2.4.5) have a unique pair of solution
(u, u1 ) in H := H01 (Ω) × L2 (Ω; Hper
1
(Y )/R), due to Lax-Milgram result. The
space H is a Hilbert space with the norm
Theorem 2.4.1. Let uε be the unique solution of (2.4.1). Then there exists
(u, u1 ) ∈ H satisfying (2.4.2) and (u, u1 ) is the unique solution of the two-
scale system (2.4.4) and (2.4.5).
Proof. Observe that (2.4.3) is the weak formulation of (2.4.4) and (2.4.5).
We introduce the bilinear form B : H × H → R defined as
Z
B[(φ, φ1 ), (ψ, ψ1 )] = A(x, y)(∇φ(x)+∇y φ1 (x, y))·(∇ψ(x)+∇y ψ1 (x, y))
Ω×Y
The coupled two-scale system (2.4.4) and (2.4.5) can be decoupled. Using
(2.4.5), one can represent u1 in terms of u. This is then used in (2.4.4) to
get the homogenized equation of u. Recall that {ej }n1 denotes the standard
basis vectors Rn . Freezing x as a parameter in (2.4.5) and substituting
Pn of ∂u
∇u(x) = j=1 ∂xj (x)ej , we get
Observe that above equation is same as (1.5.6) for each fixed x ∈ Ω. Thus,
there is a function ũ(x) such that
n
X ∂u
u1 (x, y) + (x)χj (x, y) = ũ(x).
j=1
∂x j
Note that this is the precise form we obtained in (1.5.7) except that the
coefficient depends on x, as well. Further,
Z n
!
1 X ∂χ k (x, y)
a0ik (x) = aik (x, y) − aij (x, y) dy
|Y | Y j=1
∂y j
Z
1
= A(x, y)(ek − ∇y χk ) · ei dy.
|Y | Y
CHAPTER 2. TWO-SCALE CONVERGENCE 37
2
Proof. If A is smooth, say, A ∈ C(Ω; L∞ per (Y ))
n
then, by the regularity of
j 2
χ , the function u1 (x, x/ε) ∈ L (Ω). Thus, u1 may be used as a testfunction
for the two-scale convergence. Let ξε := ∇uε − ∇u(x) − ∇y u1 x, xε . Then,
Z Z
Aε (x)ξε · ξε dx = f (x)uε (x) dx
Ω ZΩ h x i h x i
+ Aε ∇u + ∇y u1 x, · ∇u + ∇y u1 x, dx
Ω ε ε
Z h x i
− (Aε + t Aε )(x)∇uε (x) · ∇u(x) + ∇y u1 x, dx.
Ω ε
Using the coercivity of A and passing to the two-scale limit, we obtain
Z
2
α lim sup kξε k2,Ω ≤ f (x)u(x) dx
ε→0 Ω Z Z
1
− A(x, y)[∇u + ∇y u1 ] · [∇u + ∇y u1 ] dy dx
|Y | Ω Y
The term on right-hand side is zero (why!), thus completing the proof.
2.5 Summary
Chapter 3
H-Convergence
Let B(X, X ? ) denote the set of all linear bounded homomorphisms from
X to X ? . The norm on B(X, X ? ) is given as,
kAxkX ?
kAkB(X,X ? ) = sup .
x∈X kxkX
39
CHAPTER 3. H-CONVERGENCE 40
and hence,
αkxkX ≤ kAxkX ? . (3.1.3)
Step 1 (Claim: A is injective). Let Ax1 = Ax2 . Then, from (3.1.3), αkx1 −
x2 kX ≤ 0. Therefore, x1 = x2 .
kA(xn − x0 )kX ? → 0.
hg, A−1
εj
f i − hg, u? i = hg, A−1
εj
f − A−1 f + A−1
εj m
f − Lfm + Lfm − u? i.
εj m
j j j j
For large m, one can make the RHS as small as possible. Thus,
hf, A−1
εj
f i ≥ αkA−1
εj
f k2X
j j
implies that
αkLf k2X ≤ hf, Lf i
or equivalently, hL−1 u, ui ≥ αkuk2X . We set A0 = L−1 .
3.2 H-Convergence
Let Ω be an open bounded subset of Rn . Let 0 < α < β and M (α, β, Ω)
denote the set of all n × n matrices A(x) = (aij (x)) of functions such that
α|ξ|2 ≤ A(x)ξ · ξ and |A(x)ξ| ≤ β|ξ| for a.e. x ∈ Ω and for all ξ ∈ Rn .
is such that
uε * u0 weakly in H01 (Ω) and (3.2.4a)
Corollary 3.2.3 (Local Property). If {Aε } and {Bε } are two sequences in
M (α, β, Ω) which H-converges to A0 and B0 , respectively, such that Aε = Bε
in ω ⊂ Ω, for all ε, then A0 = B0 in ω.
H H
Theorem 3.2.4 (Transpose). If Aε * A0 then t Aε * t A0 .
CHAPTER 3. H-CONVERGENCE 45
Then, upto a subsequence, there is a v ∈ H01 (ω) and η ∈ [L2 (ω)]n such that
= η · ∇u0 φ dx.
ω
Thus, t Aε ∇vε ·∇uε * η ·∇u0 weak-* in D0 (Ω). Since Aε (x)∇uε (x)·∇vε (x) =
t
Aε (x)∇vε (x) · ∇uε (x), we have A0 (x)∇u0 (x) · ∇v(x) = η(x) · ∇u0 (x) a.e. in
CHAPTER 3. H-CONVERGENCE 46
and kuε kH01 (a,b) ≤ (C/α)kf kL2 (a,b) . By Eberlein-Šmuljan theorem, there exists
a subsequence of {uε }, also denoted by uε , such that
uε * u weakly in H01 (a, b),
for some u ∈ H01 (a, b). We need to find the homogenized equation which u
solves. Set ξε := aε u0ε . Note that ξε is bounded in L2 (a, b), because aε is
bounded in L∞ (a, b) and u0ε is bounded in L2 (a, b). From the equation, we
have that −ξε0 (x) = f (x) and hence ξε is bounded in H 1 (a, b). Therefore,
there exists a ξ ∈ H 1 (a, b) such that for a subsequence of ξε (denoted by
itself),
ξε * ξ weakly in H 1 (a, b).
Thus, by compact imbedding of H 1 (a, b) in L2 (a, b), ξε converges to ξ strongly
in L2 (a, b). Note that if either aε or u0ε converges strongly, then ξ is the
product of the limits of aε and u0ε . But, in general, this need not be the case.
Recall the u0ε converges to u0 weakly in L2 (a, b), therefore
1
ξε * u0 weakly in L2 (a, b).
aε
CHAPTER 3. H-CONVERGENCE 47
u is already in H01 (a, b) satisfying the boundary condition and the effective
coefficient is a0 = 1/b. The effective coefficient is bounded in L∞ (a, b) and
satisfies the ellipticity condition. Note that the choice of b depends on the
subsequence chosen and u is not unique. All the above arguments were for a
subsequences, extracted sufficiently.
However, if aε = a(x/ε) such that a is periodic. Then
Z 1
1
b= dy
0 a(y)
∂gε
∈ L2 (Ω1 ; H −1 (Ω2 ))
∂x1
because
∂gε ∂ξ 1
= ε.
∂x1 ∂x1
Thus, {gε } is bounded in H 1 (Ω1 ; H −1 (Ω2 )) and, by Aubin-Lions compactness
theorem, relatively compact in L2 (Ω1 ; H −1 (Ω2 )). Thus, gε → g strongly in
L2 (Ω1 ; H −1 (Ω2 )) and g = ξ 1 . Using the equation for ξε1 (x), we get
1 X ∂ aε1j (x1 )
ε 1 ε
ux1 (x) = ε ξ (x) − u (x) (3.2.6)
a11 (x1 ) ε j=2
∂xj aε11 (x1 )
and, for 2 ≤ i ≤ n,
This yields ξ 1 (x) = nj=1 a1j uxj . Similarly, multiplying φ ∈ Cc∞ (Ω) on both
P
sides of (3.2.7) and, passing to limit, one gets for 2 ≤ i ≤ n
i.e., wεi * φPi weakly in H01 (O). Choose φ ∈ D(O) such that φ ≡ 1 in Ω,
then
(i) φPi = Pi in Ω;
The last equation implies that there is a ηi ∈ [L2 (Ω)]n such that, for a
subsequence, t Aε ∇wεi * ηi weakly in [L2 (Ω)]n . Also, −div(ηi ) = Fi in Ω. The
fact that ∇wεi weakly converges to ei motivates to define t A0 ei = ηi . Thus,
set A0 (x) = aij (x) where aij (x) = (ηi )j . Let ω ⊂⊂ Ω be compactly contained
in Ω. Define the operator Aε : H01 (ω) → H −1 (ω) as Aε = −div(Aε ∇). By
Theorem 3.1.7, there is a A0 : H01 (ω) → H −1 (ω) such that, for a subsequence,
ε→0
hg, A−1 −1
ε f i −→ hg, A0 f i ∀f, g ∈ H
−1
(ω).
−1
For a fixed f ∈ H −1 (ω), set uε := A−1
ε f and u0 := A0 f . Then uε * u0
−1 −1
weakly in H0 (ω). Set ξε := Aε ∇(Aε ) : H (ω) → [L (ω)]n . Note that
1 2
β
kξε f k2,ω ≤ βkA−1
ε f k1,2,ω ≤ kf k−1,2,ω
α
is bounded in [L2 (ω)]n . Arguing as in the proof of Theorem 3.1.7, one can
conclude that there is a ξ0 : H −1 (ω) → [L2 (ω)]n such that
But both the LHS and RHS converge weak-* in D0 (ω). For all φ ∈ Cc∞ (ω),
using integration by parts
Z Z
i
(ξε f · ∇wε )φ dx * (ξ0 f · ei )φ dx
ω ω
and Z Z
t
(∇uε · Aε ∇wεi )φ dx * (∇u0 · t A0 ei )φ dx.
ω ω
−1
Thus, for a.e. in ω and any f ∈ H (ω),
ξ0 f = A0 ∇u0 = A0 ∇(A−1
0 f ).
Hence, ξ0 = A0 ∇(A−1
0 ).
and kuε kH01 (Ω) ≤ (1/α)kfε kH −1 (Ω) . By Eberlein-Šmuljan theorem, there exists
a subsequence of {uε }, also denoted by uε , such that (3.2.4a) is satisfied, for
some u0 ∈ H01 (Ω).
Set ξε (x) = Aε (x)∇uε (x). Note that ξε is bounded in (L2 (Ω))n , because
the entries of Aε are bounded in L∞ (Ω) and ∇uε is bounded in (L2 (Ω))n .
Therefore, there exists a ξ0 ∈ (L2 (Ω))n such that for a subsequence of ξε
(denoted by itself),
ξε * ξ weakly in (L2 (Ω))n .
Note that, in contrast, in the one-dimensional case we had the strong con-
vergence in L2 (Ω), because we had the boundedness of ξε in H 1 (Ω).
Passing to the limit, as ε → 0, in the weak formulation of (3.2.3)
Z
ξε ∇v dx = hfε , viH −1 (Ω),H 1 (Ω) , ∀v ∈ H01 (Ω),
0
Ω
R
we get Ω ξ0 ∇v dx = hf, viH −1 (Ω),H 1 (Ω) for all v ∈ H01 (Ω). Thus, −div(ξ0 ) = f
0
in Ω. Our proof is done if we show that ξ0 = A0 ∇u0 .
The compactness of H-convergence implies the existence of a matrix
2 H
A0 ∈ M (α, βα , Ω) such that, for a subsequence (still denoted by ε), Aε * A0 .
H
Further, t Aε * t A0 . Note that, as usual, ξε is a product of two weak con-
verging sequences and finding their limit is not trivial. This was cleverly
overcome in one dimensional case. For the general case, Tartar came up with
idea of using the adjoint of Aε to define some useful test functions, called the
method of oscillating test function.
For each 1 ≤ i ≤ n, let wεi ∈ H 1 (Ω) be a solution of
−div(t Aε ∇wεi ) = −div(t A0 ei ) in Ω
(3.2.8)
wεi = xi on ∂Ω.
The function wεi ∈ H 1 (Ω), for all 1 ≤ i ≤ n and satisfies the following
properties
i
wε * xi weakly in H 1 (Ω),
t
Aε ∇wεi * t A0 ei weakly in (L2 (Ω))n ,
div(t Aε ∇wεi ) converges strongly in H −1 (Ω).
CHAPTER 3. H-CONVERGENCE 52
Note that for any φ ∈ D(Ω), φwεi ∈ H01 (Ω). Thus, in particular choosing
v = φwεi in the weak formulation of (3.2.3), we get
Z
i
fε , φwε H −1 (Ω),H 1 (Ω) = ξε · ∇(φwεi ) dx
0
ZΩ Z
= ξε · (∇φ)wε dx + ξε · (∇wεi )φ dx.
i
Ω Ω
Note that the last term involves product of two weak converging sequences
in (L2 (Ω))n . To overcome this difficulty, we use φuε as a test function in
(3.2.8) to get
Z Z
t t
A0 ei ∇(φuε ) dx = Aε ∇wεi ∇(φuε ) dx
Z ZΩ ZΩ
t t t
A0 ei (∇φ)uε dx + A0 ei (∇uε )φ dx = Aε ∇wεi ∇φuε dx
Ω Ω ΩZ
+ t Aε ∇wεi ∇uε φ dx
Z Z Z Ω
t
A0 ei (∇φ)uε dx + t A0 ei (∇uε )φ dx = t
Aε ∇wεi ∇φuε dx
Ω Ω ΩZ
+ ξε · (∇wεi )φ dx
Ω
Therefore, we have
Z Z
fε , φwεi H −1 (Ω),H 1 (Ω) = ξε · (∇φ)wεi
dx + t A0 ei (∇φ)uε dx
0
ΩZ Ω Z
+ t A0 ei (∇u0 )φ dx − t A0 ei (∇φ)u0 dx
Z Z Ω Z Ω
t
ξ0 · ∇(φxi ) dx = ξ0 · (∇φ)xi dx + A0 ei (∇u0 )φ dx
Ω Z ZΩ Ω
ξ0 · ei φ dx = A0 ∇u0 · ei φ dx.
Ω Ω
CHAPTER 3. H-CONVERGENCE 53
Set wεi (x) = εwi xε . The reason behind having a factor of ε while defining
wεi is to avoid the factor of 1/ε while computing its first derivative. Thus,
h x i x 1 x
∇x wεi (x) = ∇x εwi = ε∇y wi = ∇y w i
ε ε ε ε
and the vector x x
t
Aε (x)∇wεi (x) t
= A ∇y w i
ε ε
2 n
in (L (Ω)) is also Y -periodic. Therefore, by Theorem 1.1.3, we have that
Z
t i 1 t
Aε (x)∇wε (x) * A(y)∇y wi (y) dy weakly in (L2 (Ω))n .
|Y | Y
Recall that our aim is to identify A0 such that (3.2.8) is satisfied. Consider
the function φ ∈ D(Ω) and set φε (y) = φ(εy) for y ∈ (0, 1)n and extended
to all of Rn . Using φε ∈ D(Rn ) as a test function in the cell equation of wi
above, we get
Z Z
t i t
A(y)∇y w (y) · ∇y φε (y) dy = A(y)ei · ∇y φε (y) dy
Rn Rn
CHAPTER 3. H-CONVERGENCE 54
By density, the above equality is true for all φ ∈ H01 (Ω). Therefore,
Z Z
1 i
A0 ei = A(y)∇y χ (y) dy − A(y) dy ei ,
|Y | Y Y
where χi solves the cell equation (3.2.9) where t A is replaced with A. Note
that formula is same as (1.5.7).
Note that the A0 obtained is independent of the choice of the subsequence
of Aε . Thus, u is a unique solution and all the convergences (3.2.4a) and
(3.2.4b) are true for the entire sequence and not just for subsequences (cf.
Eberlein-Šmuljan).
Recall that in the one dimensional case, we encountered the problem of
product of two weak converging sequences (recall the sequence ξε ). In the one
dimensional case, it was easy overcome this constraint by other means. How-
ever, the same idea would fail in higher dimension. The following theorem,
popular as compensated compactness, is a fix of the problem in higher dimen-
sions and is due to F. Murat and L. Tartar (cf. [Mur78a, Mur79, Tar79]).
CHAPTER 3. H-CONVERGENCE 55
Lemma 3.2.7 (div-curl lemma). Let uε and vε be two sequences in (L2 (Ω))n
such that
uε * u0 weakly in (L2 (Ω))n
H
Theorem 3.2.8 (Energy convergence). If Aε * A0 then
Z Z
Aε ∇uε .∇uε dx → A0 ∇u0 .∇u0 dx (3.2.10)
Ω Ω
where uε and u0 are, respectively, the unique solution of (3.2.3) and (3.2.5).
The energy convergence also amounts to saying that the quadratic forms
associated with the operators converge, i.e., hAε uε , uε i → hA0 u0 , u0 i. In
section §4.3 (cf. Lemma 4.1), we will observe that this is actually subject to
a special type of convergence called the Γ-convergence.
The energy functional (cf. (3.2.10)) involves a product of two weakly
converging sequences and we have shown that the limit of the product is
equal to the product of the limit. This property is not true, in general, and
was possible due to the div-curl lemma.
3.3 Correctors
We have from (3.2.4a) that
The corrector matrices are obtained by looking for functions χiε ∈ H 1 (Ω),
for 1 ≤ i ≤ n, with the following properties:
i
χε * xi weakly in H 1 (Ω),
Aε ∇χiε * A0 ei weakly in (L2 (Ω))n , (3.3.1)
div(Aε ∇χiε ) converges strongly in H −1 (Ω).
Proof. Note that {Dε } is bounded in [L2 (Ω)]n×n . Let Φ ∈ [Cc∞ (Ω)]n . Then
n
X
Φ(x) = Φi (x)ei
i=1
Therefore,
Z n Z
X Z
lim Dε Φ dx = ei Φi dx = Φ dx.
ε→0 Ω Ω Ω
i=1
rs
where t = min 2, r+s . If u0 is a solution of (3.2.5) then
Proof. Observe that if u0 ∈ Cc∞ (Ω) then we choose Φ = ∇u0 in the previous
lemma and the result is true. For any δ > 0, choose Φ ∈ Cc∞ (Ω) such that
(a) There exists a B ] (depending only on {Aε } and {Bε }) such that
t
Dε Bε Dε * B ] weak* in (D0 (Ω))n×n . (3.4.1)
(d) B ] ∈ M c, d( ab )2 , Ω .
µ(Ωε )
In the two dimensional case and when θ = µ(Ω)1 , the above result has a
nice geometric interpretation. For each θ ∈ (0, 1), consider the set of points
a1 a2
(X(θ), Y (θ)) := , a2 − (a2 − a1 )θ .
(a2 − a1 )θ + a1
Note that (X(0), Y (0)) = (a2 , a2 ) and (X(1), Y (1)) = (a1 , a1 ). The points
subscribe to a concave hyperbola H1 given by Y = a1 + a2 − a1Xa2 , for a1 ≤
X, Y ≤ a2 , above the line Y = X. The reflection of H1 along Y = X line
gives a convex hyperbola H2 with equation
a1 a2
Y = .
a1 + a2 − X
Let G(θ) := (A∗ , A∗ ) and M (θ) := (A, A) be the points on Y = X line. Then
the points G(θ), M (θ), (A∗ , A) ∈ H1 and (A, A∗ ) ∈ H2 form a square. The
condition
A∗ (x) ≤ λi (x) ≤ A(x) ∀i = 1, 2
implies that the eigenvalues of the H-limit A0 are contained in the square
describes above. The remaining two inequalities imply that the eigenvalues
are contained between two hyperbolas H3 and H4 given by the equation
1 1 1 1
H3 : + = ∗ +
X − a1 Y − a1 A − a1 A − a1
and
1 1 1 1
H4 : + = + .
a2 − X a2 − Y a2 − A∗ a2 − A
The well-known Hashin-Shtrikman (Clausius-Mossoti or Lorentz-Lorenz
or Maxwell-Garnett) inequalities is the two dimensional case with A0 is
isotropic. In this case, the inequalities reduce to
θa1 + (1 − θ)a2 + a2 θa1 + (1 − θ)a2 + a1
a1 ≤ λ ≤ a2 .
(1 − θ)a1 + θa2 + a1 (1 − θ)a1 + θa2 + a2
CHAPTER 3. H-CONVERGENCE 62
Chapter 4
Γ-Convergence
4.1 Motivation
Recall that the weak solution u of (1.2.2) can also be characterized as the
minimizer in H01 (Ω) of the functional
Z
1
J(v) = A∇v.∇v dx − hf, viH −1 (Ω),H 1 (Ω) ,
2 Ω 0
i.e.,
J(u) = min
1
J(v).
v∈H0 (Ω)
Thus, the problem of studying the asymptotic behaviour of the second order
elliptic problem
−div(Aε ∇uε ) = f in Ω
uε = 0 on ∂Ω,
with {Aε } ⊂ M (α, β, Ω) is equivalent to finding a functional J on H01 (Ω)
whose minimum is the solution of the homogenized elliptic equation such
that both the minimizers and minima of Jε converge to the minimizers and
minima of J. Thus, we need to study the convergence of functionals such
that the minimizers and minima converge.
63
CHAPTER 4. Γ-CONVERGENCE 64
Our aim, in this section, will be to generalise this basic result to an arbitrary
topological space. We recall the proof to motivate some definitions.
Theorem 4.2.1 (Extreme value theorem). Any real valued continuous func-
tion f on a closed bounded interval [a, b] attains its minimum.
Proof. We show f has a lower bound. Suppose f is not bounded below then
there exists a sequence {xn } ⊂ [a, b] such that f (xn ) > n. Since the sequence
xn is bounded, by Bolzano-Weierstrass theorem, there exists a convergent
subsequence {xnk } and let x be its limit. Thus, by continuity of f , f (xnk )
converges to f (x) which is a contradiction since f (xnk ) > nk ≥ k. Thus
there exists a infimum (greatest lower bound) C such that f (x) ≥ C for all
x ∈ [a, b]. It now remains to show that there exists a x ∈ [a, b] such that
f (x) = C.
Let {yn } ⊂ [a, b] be a sequence such that f (yn ) ≤ C + 1/n. Since C ≤
f (yn ) for all n, we have that f (yn ) converges to C (thus the sequence yn
is called minimizing sequence). By applying Bolzano-Weierstrass theorem
again, there exists a convergent subsequence {ynk } and let y be its limit.
Using continuity of f again, f (ynk ) converges to f (y). Thus f (y) = C.
Moreover, since [a, b] is closed y ∈ [a, b].
We shall, henceforth, concentrate on the minimum of the function f ,
since the corresponding result for maximum can be obtained by applying the
results to −f .
Definition 4.2.2. Let X be a topological space. A function F : X → R =
R ∪ {−∞, +∞} is said to be lower semicontinuous (lsc) at a point x ∈ X if
Note that one can, in fact, weaken the hypothesis of the extreme value
theorem.
Exercise 4.1. Prove the Extreme value theorem when f is lower semicontiu-
ous. Replace lower semicontinuity hypothesis with upper semicontinuity to
obtain the maximizer.
Exercise 4.2. Show that if F is lower semicontinuous then the sublevel set
{F ≤ α} := {x ∈ X : F (x) ≤ α} is closed for all α ∈ R.
We have already seen that the continuity property in Extreme Value
theorem can be relaxed to lower semicontinuity. Another crucial element of
the proof is the Bolzano-Weierstrass theorem which is about the compactness
of the interval [a, b].
{Ψ ≤ α} = ∩β>α Kβ
for every t ∈ (0, 1) and for every x, y ∈ X such that x 6= y, F (x) < +∞ and
F (y) < +∞.
If F is constant, then one can see that X is the set of all minimizers of F .
We now show that with strict convexity the minimizer, if exists, is unique.
(i)
F − (x) = sup lim inf inf Fn (y).
U ∈N (x) n→∞ y∈U
(ii)
F + (x) = sup lim sup inf Fn (y).
U ∈N (x) n→∞ y∈U
Γ Γ
Exercise 4.6. Show that if Fn → F , Gn → G and Fn ≤ Gn , for each n, then
F ≤ G.
Exercise 4.7. Show that if Fn Γ-converges to F , then F is lower semicontin-
uous.
Exercise 4.8. Let X be a topological vector space. Show that if Fn : X → R
is convex for each n, then Γ-lim supn Fn is convex. Also show that the Γ-
lim inf n Fn is, in general, not convex.
CHAPTER 4. Γ-CONVERGENCE 69
Therefore,
inf F + (x) ≥ lim sup inf Fn (y).
x∈U n→∞ y∈U
(i) F is coercive.
(ii) limn→∞ dn = d, where dn = inf x∈X Fn (x) and d = inf x∈X F (x). That
is, the minima converges.
Proof. Since {Fn } are equi-coercive, by Proposition 4.2.8, there is a lsc, co-
ercive function Ψ on X such that Fn ≥ Ψ. Now, by Exercise 4.6, F ≥ Ψ and
by Exercise 4.3 F is coercive.
Now, by putting U = X in Theorem 4.3.3, we get d ≥ lim supn dn . We
now need to show that d ≤ lim inf n dn . If Fn are all not identically +∞, then
lim inf n dn < +∞. Set lim inf n dn = α. By the equi-coercivity of Fn , there is
a compact set Kα such that {Fn ≤ α} ⊆ Kα , for all n. Consider,
Proof. Let {Uk }k∈N be a countable base for the topology of X. For each k,
let dnk = inf y∈Uk Fn (y). Thus, {dnk }n is a sequence in R which is compact,
hence has a subsequence {dm k }m whose limit as m → ∞ exists in R. Thus,
for each k, we have subsequence {dm k }m whose limit as m → ∞ exists in R.
CHAPTER 4. Γ-CONVERGENCE 71
Now, define F (x) = supU ∈N (x) limk→∞ inf y∈Uk Fk (y) and we have by definition
Fk Γ-converges to F .
H Γ
Example 4.1. Let Aε * A0 then we wish to show that Jε * J in the weak
topology of H01 (Ω) where
Z
Jε (u) = Aε ∇u.∇u dx
Ω
and Z
J(u) = A0 ∇u.∇u dx.
Ω
Let u ∈ H01 (Ω). We need to find a sequence {uε } in H01 (Ω) such that uε
converges to u weakly in H01 (Ω) and limε→0 Jε (uε ) = J(u). Let uε ∈ H01 (Ω)
be the solution of
−div(Aε ∇uε ) = −div(A0 ∇u). (4.3.1)
1
Then,
R R H-convergence that uε * u weakly in H0 (Ω) and
it follows from
A ∇uε .∇uε dx → Ω A0 ∇u.∇u dx. Thus, we have shown the existence of
Ω ε
a sequence {uε } converging weakly to u in H01 (Ω) such that
Now, let wε ∈ H01 (Ω) be a sequence such that wε * u weakly in H01 (Ω).
Then, the solution uε obtained in (4.3.1) minimizes the functional
Z
1
Jε (v) − A0 ∇u.∇v dx.
2 Ω
Γ
Hence Jε * J in the weak topology of H01 (Ω).
In the above example, we assume the H-convergence of the matrix coeffi-
cients to describe the Γ-limit. A general question of interest is the following:
If for any sequence of functionals, by compactness, there is a Γ-limit, then
under what conditions one can get an integral representation of Γ-limit. In
the next section, we describe the situation in one-dimension.
Proof. Let us assume f (x, ·) ∈ C 1 (R) for all x ∈ (a, b). Due to the growth
conditions and continuity of f ,
Hence,
Z b
?
Φ(x)u0 (x) − f (x, u0 (x) dx
F (φ) =
a
Z b
= f ? (x, Φ(x) dx
a
Z b Z x
?
= f x, − φ(t) dt dx
a a
where ρε are the sequence of mollifiers. Observe that fε are convex in the
second variable and, by Jensen’s inequality, fε ≥ f . Also, observe that
limε fε? (x, ξ ? ) = f ? (x, ξ ? ) for all x ∈ (a, b) and ξ ? ∈ R. We have, for each ε,
Z b Z x
? ?
Fε (φ) = fε x, − φ(t) dt dx ∀φ ∈ L1 (a, b).
a a
Proof. Let v ∈ C([a, b]) and φ ∈ L1 (a, b). Also, let (xi−1 , xi ) be k number of
partitions of (a, b) for i = 1, 2, . . . , k such that x0 = a and xk = b. Consider,
Z b k
X Z xi
(gn (x, v) − g(x, v)) φ dx ≤ [gn (x, v) − gn (x, v(xi ))] φ dx
a i=1 xi−1
k
X Z xi
+ [gn (x, v(xi )) − g(x, v(xi ))] φ dx
i=1 xi−1
Xk Z xi
+ [g(x, v(xi )) − g(x, v(x))] φ dx
i=1 xi−1
Lemma 4.4.3. Let gn : Ω×Rn → [0, +∞) satisfy hypotheses H1 and H2, for
all n. Then, there exists a subsequence of {gn } and a g : (a, b)×R → [0, +∞)
such that gn (·, ξ) weak* converges to g(·, ξ) for all ξ ∈ R.
The proof of above lemma and theorem are being skipped and can be
found in [Bra02].
Example 4.2. Let 0 < α ≤ aε (x) ≤ β < +∞ and g ∈ L2 (a, b). Let Fε :
H01 (a, b) → R be defined as
Z b
1 0 2
Fε (u) = aε (x)|u | − gu dx.
a 2
ξ2
Now, set fε (x, ξ) := aε (x)|ξ|2 . Then, fε? (x, ξ ? ) = 4aε (x)
. But, for each ξ ? ∈
ξ2
Rn , fε? (·, ξ ? ) converges weak* in L∞ (a, b) to f ? (·, ξ ), where f ? (x, ξ ? ) =
?
4b(x)
and
1 1
* .
aε (x) b(x)
Chapter 5
Bloch-Floquet Homogenization
77
CHAPTER 5. BLOCH-FLOQUET HOMOGENIZATION 78
= |ξ|2 û(ξ).
More generally, any m-th order linear differential equation with constant
α
P
coefficients P (D)u = f where P (D) = |α|≤m aα D will transform in to
an algebraic eqaution P (ıξ)û(ξ) = fˆ(ξ). The Laplacian is a particular case
of the elliptic operator −∆ + c(x) with c ≡ 0. For c(x) 6= 0 (without loss
of generality assume c(x) ≥ 0), the Bloch theorem gives the generalised
eigenfunction for −∆ + c(x) when c is Y -periodic, for any given reference cell
Y ⊂ Rn .
Then U (p)L = LU (p). In fact, U (p)e−ıLs = e−ıLs U (p). Let us first consider
the one dimension situation with c ∈ Cc∞ (R) with bounded derivatives and
L : S(R) → S(R) defined as
d2
L := − + c(x).
dx2
If c is 2π-periodic and, hence, c admits a uniformly convergent Fourier series
X
c(x) = cm eımx
m∈Z
where Z π
1
cm = c(x)e−ımx dx.
2π −π
If u ∈ S(R) then
Z
1
\
Lu(x)(ξ) = ξ 2 û(ξ) + √ c(x)u(x)e−ıξx dx
2π R
Z !
1 X
= ξ 2 û(ξ) + √ cm eımx u(x)e−ıξx dx
2π R m∈Z
Z
X 1
2
= ξ û(ξ) + cm √ u(x)e−ı(ξ−m)x dx
m∈Z
2π R
X
= ξ 2 û(ξ) + cm û(ξ − m).
m∈Z
Thus, Lu(ξ)
c depends only on the values û(ξ − m) for all m ∈ Z. But recall
that û(ξ − m) = eıxm
\ u(x)(ξ). This suggests that the operator L depends on
the “modulation” by all m ∈ Z.
Define T : L2 (R) → H by
[(T f )(η)]k = fˆ(η + k).
d 2
2
For L = − dx 2 + c(x) on L (R),
Z ⊕
−1
T LT = Lη dη.
(− 12 , 21 ]
CHAPTER 5. BLOCH-FLOQUET HOMOGENIZATION 81
( by Fubini’s Theorem)
Z 2π X ! Z
2
= |f (x + 2πm)| dx = |f (x)|2 dx.
0 m∈Z R
Thus, T is well defined and admits a unique isometry extension. To see that
T is onto H, we compute T ? . For any g ∈ H, x ∈ [0, 2π] and m ∈ Z
Z 1
2
?
(T g)(x + 2πm) = eı2πmη gη (x) dη.
− 12
CHAPTER 5. BLOCH-FLOQUET HOMOGENIZATION 82
Further,
Z
?
kT gk22 = |(T ? g)(y)|2 dy
R
!
Z2π X
= |(T ? g)(2πm + x)|2 dx
0 m∈Z
2
!
Z 2π X Z 2π
= eı2πmη gη (x) dθ dx
0 m∈Z 0
Z 2π Z 2π
2
= |gη (x)| dθ dx (Plancherel’s identity)
0 0
= kgk2 .
2
Similarly, (T f )0η (2π) = eı2πη (T fη )0 (0). Thus, for each η, (T f )η ∈ D((− dx
d
2 )η )
and
d2
− 2 (T f ) = U (−f 00 )η .
dx η
Then Z ⊕
d2
−1
T − 2 +c T = Lη dη.
dx (− 12 , 21 ]
Proof. Let c be the η-independent operator acting on the fiber H = L2 [0, 2π)
by (cη f )(x) = c(x)f (x) for 0 ≤ x ≤ 2π. It is sufficient to prove that
Z ⊕
−1
T cT = cη dη.
(− 12 , 12 ]
For f ∈ S(R),
X
(T cf )η (x) = e−ı2πmη c(x + 2πm)f (x + 2πm)
m∈Z
X
= c(x) e−ı2πmη f (x + 2πm)
m∈Z
= cη (T f )η (x).
Thus, we observe that the above space consists of (η, Y )-Bloch Periodic func-
tions. For any irrational η can be approximated by rationals by varying m
and noting that the sets of roots of unity is dense in S 1 .
CHAPTER 5. BLOCH-FLOQUET HOMOGENIZATION 85
The sum is well defined because it has finite number of terms because f has
compact support. Note that fb (y; η) is Y -periodic in y variable because
X
fb (y + 2π; η) := f (y + 2πp)e−ı(y+2πp)·η = fb (y; η).
p+1∈Zn
In the above relation we have used the fact that eı2πp·k = 1. Observe that
X
eıy·η fb (y; η) = f (y + 2πp)e−2ıπp·η .
p∈Zn
CHAPTER 5. BLOCH-FLOQUET HOMOGENIZATION 86
Thus,
Z X Z
ıy·η
e fb (y; η) dη = f (y) + f (y + 2πp) e−2ıπp·η dy
Y ? Y ?
p∈Zn
p6=0
e−ıπp − eıπp
X
= f (y) − f (y + 2πp) dy
p∈Zn
2ıπp1 . . . pn
p6=0
= f (y).
Therefore, we have proved the results for all functions in D(Rn ). Similarly,
one can prove the Parseval’s formula for functions in D(Rn ). The Bloch
transform is a linear map on D(Rn ) bounded on L2 (Rn ). Define B : D(Rn ) →
L2 (Y × Y ? ) as Bf = fb . B is a bounded operator w.r.t L2 -norm. Consider
Z Z X
2
kBf k2 = kfb k2 ≤ 2
|f (y + 2πp)e−ı(y+2πp)·η |2 dη dy
Y Y ? p∈Zn
Z X Z
2 −ı(y+2πp)·η 2
= |f (y + 2πp)| |e | dη dy
Y p∈Zn Y?
XZ
?
= |Y | |f (y + 2πp)|2 dy
p∈Zn Y
Z
= |f (y)|2 dy = kf k22 .
Rn
and extended unitarily to to L2 (Rn ). Further, fb (y; η) = e−ıy·η fz (y; η) for all
f ∈ D(Rn ).
The following theorem explains the sense in which the Bloch transform
leaves the periodic functions invariant.
Theorem 5.4.4 (Invariance of Periodic Functions). Let Y = [0, 2π)n and
c : Y → C be such that c ∈ L∞ (Y ) extended Y -periodically to Rn . For any
f ∈ L2 (Rn ), (cf )b (y; η) = c(y)fb (y; η).
Proof. It is enough to prove the result for f ∈ D(Rn ). Consider
X
(cf )b (y; η) = c(y + 2πp)f (y + 2πp)e−ı(y+2πp)·η
p∈Zn
X
= c(y) f (y + 2πp)e−ı(y+2πp)·η
p∈Zn
= c(y)fb (y; η).
where {λm } are positive real eigenvalues and {φm (y)} are the corresponding
eigenvectors, for each m ∈ N, such that {φm } form an orthonormal basis
of L2per (Y ) and 0 ≤ λ1 ≤ λ2 ≤ . . . diverges and each eigenvalue has finite
multiplicity.
where, {φm } are the eigenfunctions corresponding to the shifted operator A(η)
and fbm (η), for each η ∈ Y ? , is the m-th Bloch coefficient of f defined as
Z
m
fb (η) := f (y)e−ıy·η φm (y; η) dy.
Rn
Proof. It is enough to prove the result for f ∈ D(Rn ). Recall that, for each
η ∈ Y ? , fb (·; η) ∈ L2per (Y ). Hence, by spectral decomposition of A(η),
∞
X
fb (y; η) = fbm (η)φm (y; η),
m=1
where Z
fbm (η) = fb (y; η)φm (y; η) dy.
Y
But,
Z X
fbm (η) = f (y + 2πp)e−ı(y+2πp)·η φm (y; η) dy
Y p∈Zn
Z X
= f (y + 2πp)e−ı(y+2πp)·η φm (y + 2πp); η) dy
Y p∈Zn
Z
= f (y)e−ıy·η φm (y; η) dy.
Rn
CHAPTER 5. BLOCH-FLOQUET HOMOGENIZATION 90
um m
b (η)λm (η) = fb (η)
fbm (η)
um
b (η) = .
λm (η)
Set ψm (y; η) := {eıy·η φm (y; η)}. Then, for each η ∈ Y ? , ψm (·; η) forms a
basis of L2 (Rn ). Thus, L2 (Rn ) can be identified with L2 (Y ? ; `2 (N)). Let us
compute ψ(y + 2π`):
where
Z Z
∂v ∂v
R= ajk (y) (ıηj − ıηj0 )v dy + ajk (y)(ıηk − ıηk0 )v dy
Y ∂yk ∂yj
ZY
+ ajk (y)(ηk ηj − ηk0 ηj0 )|v|2 dy.
Y
By Cauchy-Schwarz’s inequality,
Z
0
|R| ≤ C0 |η − η | (|∇v|2 + |v|2 ) dy.
Y
By min-max principle,
a(v, v; η)
λm (η) = min max
1
W ⊂Hper (Y ) v∈W kvk22,Y
1
where W is a m-dimensional subspace of Hper (Y ). Using the estimate on R,
we deduce that
λm (η) ≤ λm (η 0 ) + C0 |η − η 0 |
for a suitable constant C0 . Interchanging η and η 0 , we obtain
|λm (η) − λm (η 0 )| ≤ C0 |η − η 0 |.
There exists a solution to the above equation which is unique upto an additive
multiple of φm . Hence, the RHS satisfies the compatibility condition or
Fredhölm alternative. Therefore,
∂ 2 φm
Z
Tm (η) φ dy = 0
Y ∂ηj ∂ηk m
yields a formula for the Hessian matrix Dη2 λm (η m ) in terms of φm . Thus,
1 ∂ 2 λm
1 ∂A(η) ∂λm ∂φm
(η) = hajk φm , φm i + − , φm
2 ∂ηj ∂ηk 2 ∂ηj ∂ηj ∂ηk
1 ∂A(η) ∂λm ∂φm
+ − , φm .
2 ∂ηk ∂ηk ∂ηj
Let us summarise the properties of the eigenvalues λm (η) and eigenvectors
φm (y; η).
Theorem 5.4.13. The origin is a critical point of the first Bloch eigenvalue,
i.e., ∂λ
∂ηj
1
(0) = 0 for all j = 1, ..., n.Further, the Hessian of λ1 at η = 0 is
given by
1 ∂ 2 λ1
(0) = a0jk ∀j, k = 1, ..., n.
2 ∂ηj ∂ηk
The derivatives of the first Bloch mode can also be calculated and they are as
follows:
∂φ1 1
(y; 0) = ı|Y |− 2 wj (y) ∀j = 1, ..., n.
∂ηj
1
Proof. Use the information λ1 (0) = 0 and φ1 (y; 0) = |Y |− 2 in the Taylor
expansion with η = 0.
where, for each m ∈ N, ε > 0 and ξ ∈ ε−1 Y ? , the m-th Bloch coefficient of f
is Z
m,ε −n/2
fb (ξ) = ε f (x)e−ıx·ξ φεm (x; ξ) dx.
Rn
CHAPTER 5. BLOCH-FLOQUET HOMOGENIZATION 95
Setting
1 ∂ 2 λ1
a0jk = (0)
2 ∂ηj ηk
2u
Then nj,k=1 a0jk ξk ξj uˆ0 (ξ) = fˆ(ξ) and A0 u0 := − nj,k=1 a0jk ∂x∂ j ∂x
P P 0
k
= f (x).
The only flaw in the above argument is that in passing to limit we have
not checked uniform compact support of the sequence. To overcome this
difficulty we use cut-off function technique to localize the equation.
CHAPTER 5. BLOCH-FLOQUET HOMOGENIZATION 97
Proof. The first Bloch transform gb1,ε (ξ), a priori defined for
ε−1 ε−1 n
ξ ∈ ε−1 Y ? = (− , )
2 2
Z 21
x x
kgε k2,Rn |φ1 ( ; εξ) − φ1 ( ; 0)|2 dx ≤ C0 kφ1 (y; εξ) − φ1 (y; 0)k2,Y .
K ε ε
By Lipschitz continuity of η 7→ φ1 (·, η), the second term in the right side is
bounded above by C0 εξ. Thus, if |ξ| ≤ M then it is bounded above by cM ε
and so, in particular, it converges to zero in L∞ n
loc (R ).
Aε (φuε ) = φf + gε + hε in Rn ,
CHAPTER 5. BLOCH-FLOQUET HOMOGENIZATION 98
where
n n
X ∂φ X ∂ 2φ
gε = −2 σjε − aεjk uε ,
j=1
∂xj j,k=1 ∂xj ∂xk
n
X ∂uε
σjε (x) = aεjk ,
k=1
∂xk
n
X ∂aεjk ∂φ
hε = − uε .
j,k=1
∂x j ∂xk
Using the arguments given in the remark above, we can pass to the limit
above, since φuε is bounded in H 1 (Rn ). Neglecting all the harmonics cor-
responding to m ≥ 2 and considering only the m = 1 yields at the limit
n
1 X ∂ 2 λ1 [ n 1,ε n 1,ε
(0)ξj ξk (φu0 )(ξ) = (φf )(ξ) + lim ε 2 gb (ξ) + lim ε 2 ĥb (ξ).
d
2 j,k=1 ∂ηj ∂ηk ε→0 ε→0
(5.5.1)
The sequence σjε is bounded in L2 (Ω). Therefore, we can extract a subse-
quence (still denoted by ε) which is weakly convergent in L2 (Ω). Let σj0
denote its limit and its extension by zero outside Ω. Using this convergence
and the definition of gε , we see that
n n
X ∂φ X ∂ 2φ
gε * g0 := −2 σj0 − M(ajk ) u0 weakly in L2 (Rn ),
j=1
∂xj j,k=1 ∂xj ∂xk
strongly in L∞ n
loc (R ). On the other hand, after integraing by parts, the first
term in the RHS of the decomposition of εn/2 h1,ε b becomes
n Z 2
∂φ ∂uε
X ∂ φ ∂φ x
ε
ajk uε + − ıξj uε e−ıx·ξ φ1 ; 0 dx.
j,k=1 Rn ∂x j ∂xk ∂x k ∂xj ∂x k ε
1
Choosing φ1 (y; 0) = |Y |− 2 , it is easily seen that the above integral converges
weakly in L2 (Rn ) to
n Z
∂ 2φ
− 12
X ∂φ
|Y | M(ajk ) u0 − ıξj M(ajk ) u0 e−ıx·ξ dx
j,k=1 R
n ∂x j ∂x k ∂x k
n Z
− 12
X ∂φ −ıx·ξ
+ |Y | σk0 e dx.
k=1 R
n ∂xk
Using this information in (5.5.1) and using Theorem 5.4.13, we conclude that
n n Z
X
0 − 12
X ∂φ −ıx·ξ
[
ajk ξj ξk (φu0 )(ξ) = (φf )(ξ) − |Y |
d σk0 e dx
j,k=1 k=1 R
n ∂xk
n Z
X
− 12 0 ∂φ
−ı ξj |Y | ajk u0 e−ıx·ξ dx.
j,k=1 Rn ∂x k
This is the localized homogenized equation in the Fourier space. Taking in-
verse Fourier transform of the above equation, we obtain
n n
X ∂φ X
0 ∂ ∂φ
A0 (φu0 ) = φf − σk0 − a u0 in Rn .
k=1
∂xk j,k=1 jk ∂xj ∂xk
Since the above relation is true for all φ in D(Ω), the desired conclusions
follow. In fact, let us choose φ(x) = φ0 (x)eımx·ν , where ν is a unit vector
in Rn and φ0 (x) ∈ D(Ω) is fixed. Letting m → ∞ in the resuling relation
and varying the unit vector ν, we can easily deduce, successively, that σj0 =
Pn 0 ∂u0
k=1 ajk ∂xk in Ω and A0 u0 = f in Ω.
Appendices
101
Bibliography
[BM] J. Ball and F. Murat, W 1,p quasi convexity and variational prob-
lems for multiple integrals, Journal of Functional Analysis 58.
103
BIBLIOGRAPHY 104
convergence sequence
H-, 43 minimizing, 65
energy, 55 Spagnolo, ii
correctors, 55
Tartar, ii, 54
De Giorgi, iii two-scale
strong convergence, 31
energy two-scale converge, 24
convergence, 55 two-scale convergence, 19, 24
functional, 55
Vanninathan, 59
function
convex, 67
strictly convex, 67
homogenization, i
periodic, 1
reiterated, 26
Kesavan, 59
Lax-Milgram theorem, 3
lemma
div-curl, 55
limit
H-, 44
multi-scale, 26
Murat, ii, 54
Rajesh, 59
109
Periodic test functions are significant in homogenization problems as they allow for capturing the oscillatory behavior of coefficients within differential equations. Such functions help in identifying the two-scale structure of solutions, treating variables at both the microscopic and macroscopic levels simultaneously . The use of periodic test functions facilitates the formulation of two-scale asymptotic expansions, enabling the effective description of phenomena over heterogeneous media by revealing periodic properties in the solutions . By using these functions, one can model how composite materials with periodic structures behave at different scales, thus aiding in the prediction of effective macroscopic properties from microscopic details ."}
The density of function spaces plays a significant role in achieving convergence results within homogenization theory. Two-scale convergence, a concept introduced by G. Nguetseng and further developed by G. Allaire, is instrumental in capturing the oscillatory behavior of functions in homogenization problems . It connects weak convergence in Lp spaces to a form of generalized convergence that considers both macroscopic (x) and microscopic (x/ε) scales . The density of certain function spaces, such as those involving Y-periodic functions in Lp loc(Rn), allows derivations of effective properties of materials and provides the mathematical foundation for two-scale asymptotic expansions . This framework overcomes the limitations of classical convergence methods by characterizing the simultaneous behavior on different spatial scales, which is crucial for accurately describing the macroscopic behavior of composite materials in terms of their microscopic properties . Therefore, density ensures the completeness and readiness of function spaces to support the nuanced requirements of two-scale limits in homogenization.
Unitary extension of the Bloch and Zak transforms to L2 spaces is necessary for proper spectral analysis of periodic operators, as it ensures completeness and orthogonal decomposability in the spectral study of periodic functions. By doing so, L2 per spaces can be used to represent functions across different periodic cells and frequencies, which is crucial for analyzing their spectral properties in terms of Bloch waves and eigenvalues . This extension allows transforming spectral problems into a sequence of problems that are more manageable and helps retain inner product properties via the Plancherel identity, essential for maintaining the energy spaces' structure . Furthermore, this unitary transformation supports homogenization, demonstrating how microscopic periodic structures relate to their macroscopic properties, which is imperative for understanding material behaviors in physical applications .
The connection between the Bloch spectrum and physical properties of materials in periodic structures is primarily established through the Bloch-Floquet theory. Bloch waves describe the behavior of electrons in periodic potentials by allowing the characterization of eigenvalues (Bloch eigenvalues) and eigenfunctions (Bloch waves) associated with the symmetrical properties of the structure. This is important for understanding how waves propagate in the medium, thus correlating the spectrum to the macroscopic properties of materials. The periodic nature of the material gives rise to a band structure described by the Bloch eigenvalues, which in turn influences the material's physical properties, such as electrical conductivity and optical behavior . Homogenization aids in relating these micro-level periodic interactions to macro-level properties by effectively averaging the properties over many unit cells . The process of studying Bloch waves in this context helps in deriving generalized macroscopic behavior from microscopic interactions .
Y-periodicity in homogenization allows the mathematical modeling of materials with periodic structures, aiding in solving differential equations with rapidly oscillating coefficients by transforming them into an equivalent problem with homogenized coefficients. This periodic framework assumes heterogeneities are small and evenly distributed, which is practical for many applications . By treating the material as Y-periodic, it becomes possible to apply techniques such as the two-scale asymptotic expansion, where solutions can be expressed as expansions in terms of a small parameter, capturing both macroscopic and microscopic behavior . This approach simplifies the computational process by averaging out the heterogeneities of the material, effectively replacing it with a homogeneous one that maintains the essential features of the original heterogeneous material . This methodology is critical for addressing the challenges in computing solutions on scales smaller than the material's structure, ensuring accuracy and feasibility in practical computations ."}
The Bloch decomposition theorem reveals that every function in L2(Rn) can be uniquely decomposed into a continuum of components indexed by the Bloch wavevectors η, with each component defined on a periodic cell Y, corresponding to the periodic structure. This decomposition indicates that periodic operators have spectra consisting of bands, and the spectral properties such as band gaps are directly derived from how these decompositions interact across the periodic domain and the space of Bloch wavevectors .
Two-scale convergence is a method that is effective in identifying the oscillatory nature of sequences because it captures details that weak convergence misses. While weak convergence in Lp(Ω) averages out the effects of oscillations, failing to fully describe the oscillation mechanisms, two-scale convergence generalizes weak convergence by incorporating periodic test functions that match the oscillation scales . This allows for a more detailed capturing of these oscillations by using test functions that reflect the order of oscillations present in the sequence . Moreover, weak convergence might not detect fine structures when sequences are inherently oscillatory, as demonstrated in cases where weak limits fail to capture the internal oscillations, which two-scale convergence can reveal, providing limits that include such periodic behaviors . Therefore, two-scale convergence offers a more robust framework for analyzing sequences with oscillatory components that would otherwise be overlooked by traditional weak convergence methods .
The compactness theorem helps analyze bounded sequences in Lp(Ω) by ensuring subsequential convergence in two-scale convergence. For any bounded sequence {uε} in Lp(Ω), the theorem guarantees the existence of a subsequence which converges two-scale, i.e., uε 2s ⇀ u for some function u in Lp(Ω × Y). This property is crucial because two-scale convergence, stronger than weak convergence but weaker than strong convergence, captures both the macroscopic limit and the microscopic oscillations of the sequence. It provides a framework for studying homogenization problems where sequences possess inherent oscillatory behavior due to small-scale structures . Moreover, the boundedness in Lp implies weak convergence, which is necessary for applying the compactness theorem, thus integrating bounded sequences within a broader analysis scheme that includes oscillatory effects .
Young's inequality helps in integrating product terms by providing bounds on convolution integrals, which is useful in Lp spaces as it allows us to handle functions with different exponents and still maintain integrability . Jensen's inequality is applied to convex functions to bound an integral by evaluating the function at the integral of its argument, thereby controlling integrals involving convex transformations of Lp functions . Together, these inequalities are instrumental in homogenization problems by simplifying complex integral expressions that arise in the study of oscillatory or periodic functions in Lp spaces, thus facilitating the analysis of their limits and convergence .
The example with the function \( u(x, y) = \sin(2\pi y) \) illustrates that two-scale convergence captures oscillations that are not reflected in strong convergence. In this case, the sequence \( u_\varepsilon(x) = \sin\left(\frac{2\pi x}{\varepsilon}\right) \) converges weakly to 0 in \( L^2([0, 1]) \) because the oscillations average out. However, it does not strongly converge since the norm \( \| \sin(2\pi x / \varepsilon) \|_{L^2([0, 1])} = 1/2 \) remains constant . Despite the lack of strong convergence, the two-scale limit for \( u_\varepsilon(x) \) is \( u(x, y) = \sin(2\pi y) \) on \([0, 1] \times [0, 1]\). This demonstrates that two-scale convergence can represent oscillatory patterns through additional variables, whereas strong convergence cannot capture such periodic fluctuations .