+ All Categories
Home > Documents > Notes on Sobolev Spaces - NTNU on Sobolev Spaces Peter Lindqvist Norwegian University of Science and...

Notes on Sobolev Spaces - NTNU on Sobolev Spaces Peter Lindqvist Norwegian University of Science and...

Date post: 18-May-2018
Category:
Upload: truongtuong
View: 223 times
Download: 2 times
Share this document with a friend
28
Notes on Sobolev Spaces Peter Lindqvist Norwegian University of Science and Technology 1 L p -SPACES 1.1 Inequalities For any measurable function u : A [−∞, ], A R n , we define up = up, A = A |u( x)| p dx 1 p and, if this quantity is finite, we say that u L p (A). In most cases of interest p 1. For p = we set u= u, A = ess sup xA |u( x)|. The essential supremum is the smallest number M such that |u( x)|≤ M for a.e. x A. For example 1 , if u :[a, b] R n , is continuous, then it is easy to see that b a |u( x)| p dx 1 p −→ p→∞ max axb |u( x)|. The fundamental inequalities uv1 ≤‖up vq , 1 p + 1 q = 1, H ¨ OLDER u + vp ≤‖up + vp , MINKOWSKI 1 An interesting application of this fact in connection with the heat equation u t = u xx is given in [BB, Ch.24, pp 145-150]. 1
Transcript

Notes on Sobolev Spaces

Peter Lindqvist

Norwegian University of Science and Technology

1 Lp-SPACES

1.1 Inequalities

For any measurable function u : A→ [−∞,∞], A ∈ Rn, we define

‖u‖p = ‖u‖p, A =

A

|u(x)|p dx

1p

and, if this quantity is finite, we say that u ∈ Lp(A). In most cases of interest p ≥ 1.

For p = ∞ we set

‖u‖∞ = ‖u‖∞,A = ess supx∈A

|u(x)|.

The essential supremum is the smallest number M such that |u(x)| ≤ M for a.e.

x ∈ A.

For example1, if u : [a, b]→ Rn, is continuous, then it is easy to see that

b∫

a

|u(x)|p dx

1p

−→p→∞

maxa≤x≤b

|u(x)|.

The fundamental inequalities

‖uv‖1 ≤ ‖u‖p‖v‖q,1

p+

1

q= 1, HOLDER

‖u + v‖p ≤ ‖u‖p + ‖v‖p, MINKOWSKI

1An interesting application of this fact in connection with the heat equation ut = uxx is given

in [BB, Ch.24, pp 145-150].

1

where p ≥ 1, can be derived in many ways. For example, taking the “elementary

inequality”

ab ≤ap

p+

bq

q(a, b ≥ 0, p + q = pq) YOUNG

for granted, we obtain the Holder inequality (choose a = u(x)/‖u‖p, b = v(x)/‖u‖qand integrate the resulting inequality). The Holder inequality implies the Minkowski

inequality.

Remarks:

1) The special case

A

|uv| dx ≤√√∫

A

u2 dx

√√∫

A

v2 dx CAUCHY

is called the Cauchy inequality.

2) If 1 > p > 0, then the Minkowski inequality is reversed for positive func-

tions! That is why one usually has p ≥ 1.

3) If mes (A) < ∞, then the Holder inequality shows that Lp1 (A) ⊂ Lp2 (A), if

p1 ≥ p2. This is not true in general, if mes (A) = ∞. (Find a simple ex.!)

However, if 1 ≤ p1 < p < p2 ≤ ∞, then the Holder inequality implies that

‖u‖p,A ≤ ‖u‖λp1,A‖u‖1−λp2,A

where1

p=λ

p1

+1 − λ

p2

.

That is, if u ∈ Lp1 ∩ Lp2 , then u ∈ Lp for all intermediate p.

4) Suppose that 0 < mes (A) < ∞. The function

φ(p) =

1

mes (A)

A

|u(x)|p dx

1p

, 0 < |p| < ∞,

φ(0) = exp

1

mes (A)

A

log |u(x)| dx

,

φ(−∞) = ess infA|u(x)|, φ(+∞) = ess sup

A

|u(x)|

2

is increasing2 as −∞ ≦ p ≦ +∞.

Let us finally mention that

Λ

1

mes (A)

A

u(x) dx

1

mes (A)

A

Λ(u(x)) dx JENSEN

whenever Λ : R→ R is a convex function (0 < mes (A) < ∞, u : A→ [−∞,∞]

is measurable).

Example:

e

1∫

0

u(x) dx

≤1∫

0

eu(x) dx

1

mes (A)

A

|u(x)| dx

p

≤ 1

mes (A)

A

|u(x)|p dx

Remark: Discrete versions of the above inequalities are

∣∣∣∣∣∣∣

k

akbk

∣∣∣∣∣∣∣

≤ p

√∑

k

|ak|p q

√∑

k

|bk|q;

n a1a2 . . . an ≤ |a1|n + |a2|n + . . . + |an|n

orn√

q1q2 . . . qn︸ ︷︷ ︸

GEOMETRIC MEAN

≤q1 + q2 + . . . + qn

n︸ ︷︷ ︸

ARITHMETIC MEAN

(qk ≥ 0);

p

√∑

k

|ak + bk|p ≤ p

√∑

k

|ak|p + p

√∑

k

|ak|p.

Especially, limp→∞

p√∑

k |ak|p = supk

|ak|.

Example: Draw the curves p√

|x|p + |y|p = 1 (0 < p ≦ ∞) in the xy-plane.

2J. Moser’s celebrated method to relate the maximum and the minimum of the solution to a

partial differential equations is based on this.

3

For p = 2 we have the parallelogram3 law

‖u + v‖22 + ‖u − v‖22 = 2‖u‖22 + 2‖v‖22.

Remark: So-called reverse Holder inequalities like

1

mes (Q)

Q

|u|

1p

≤C

mes (Q)

Q

|u|

valid for every cube with some fixed p > 1 (!) play a central role for solutions u to certain

partial differential equations. (Such an inequality cannot hold for arbitrary functions.)

The method is due to F. W. GEHRING.

1.2 (Strong) convergence in Lp

Let A ∈ Rn and fix p ≥ 1. Then ‖ ‖p defines a semi-norm in the vector space Lp(A),

i.e.,

(i)’ 0 ≤ ‖u‖p < ∞, u ∈ Lp(A)

(ii) ‖λu‖p = |λ|‖u‖p, u ∈ Lp(A), −∞ < λ < ∞

(iii) ‖u + v‖p ≤ ‖u‖p + ‖v‖p, u, v ∈ Lp(A).

However, this is, strictly speaking, not a norm, the reason being that ‖u‖p,A =0 ⇐⇒ u(x) = 0 for a.e. x ∈ A. We agree to say that u = v in Lp(A), if u(x) = v(x)

for a.e. x ∈ A. With this convention ‖ ‖p is a norm, i.e. (i)’ can be replaced by

(i) 0 ≤ ‖u‖p < ∞, and ‖u‖p = 0⇐⇒ u = 0 in Lp.

(Strictly speaking, this Lp-space consists of equivalence classes of functions, but

here there is no point in maintaining this distinction.)

Theorem 1 (RIESZ-FISCHER) The Lp-spaces, p ≥ 1, are Banach spaces. That

is, if u1, u2, . . . is a Cauchy sequence in Lp(A), then there is a function u ∈ Lp(A)

such that ‖uk − u‖p,A → 0, as k →∞.3The counterpart for general exponents p is more involved. A pair of inequalities (CLARK-

SON’s inequalities) will replace the parallelogram law [A, p.37].

4

Proof: Suppose that u1, u2, . . . is a Cauchy sequence in Lp(A), i.e., given ε > 0 there

is an index Nε such that ‖uk − u j‖p < ε, when k, j ≥ Nε. Let us consider the case p , ∞(the case p = ∞ is simpler, and its proof will be skipped). We can construct indices

1 ≤ n1 < n2 < . . . such that

‖unk− unk+1

‖p <1

2k(k = 1, 2, 3, . . .).

Define

gN(x) =

N∑

k=1

|unk− unk+1

|, g(x) = limN→∞

gN(x) =

∞∑

k=1

|unk− unk+1

|

Then

0 ≤ g1(x) ≤ g2(x) ≤ . . . ≤ g(x) ≦ ∞

and ∫

A

|g|p =∫

A

∣∣∣∣∣

limN→∞

gN

∣∣∣∣∣

p

=

A

limN→∞|gN |p ≤ lim

N→∞

A

|gN |p

by Fatou’s lemma. By the construction ‖gN‖p ≤ 12+

14+. . .+ 1

2N < 1 and so∫

A

|g|p ≤ 1. Thus

g(x) < +∞ for a.e. x ∈ A. This means that the series |un2(x)−un1

(x)|+ |un3(x)−un2

(x)|+ . . .converges for a.e. x ∈ A. So does a fortiori, the series

un2(x) +

∞∑

k=1

(unk+1− unk

).

The partial sums of this series are plainly un2(x), un3

(x), un4(x), . . . Hence

u(x) = limk→∞

unk(x)

exists and is finite for a.e. x ∈ A (If you insist on having an everywhere defined function,

set u(x)=0 in a subset of measure zero.)

By Fatou’s lemma (again!)

A

|u(x) − u j(x)|p dx =

A

(

limk→∞|unk

(x) − u j(x)|p)

dx ≤ limk→∞

A

|unk(x) − u j(x)|p dx ≤ εp

whenever j > Nε. This shows that ‖u − u j‖p,A → 0 as j→ ∞.

The proof yields information about pointwise behaviour.

Corollary 2 Suppose that uk → u in Lp(A), i.e., ‖uk − u‖p,A → 0. Then there is a

subsequence that converges a.e. in A : u(x) = limk→∞

unk(x) for a.e. x ∈ A.

5

1.3 The dual of Lp

If X is any Banach space, its dual X∗ is the collection of all continuous linear

functions (functionals) l : X → R . The norm of l is defined as

‖l‖X∗ = supx∈X

(

|l(x)|‖x‖X

)

= sup‖x‖X≤1

|l(x)|

and the continuity is equivalent to ‖l‖X∗ < ∞. Note that

|l(x)| ≤ ‖l‖X∗‖x‖.

Let 1 ≤ p < ∞ and fix a function g ∈ Lq(A). Then

l( f ) =

A

f (x)g(x) dx ( f ∈ Lp(A))

is well-defined and linear. By Holder’s inequality |l( f )| ≤ ‖ f ‖p‖g‖q. Hence ‖l‖∗ ≤‖g‖q. Here equality is attained for the choice f = |q|q−2g. (The case q = ∞ is

slightly different.) Thus

‖l‖∗ = ‖g‖q.The essential fact here is that virtually all continuous functionals in Lp (p , ∞)

are of this form!

Theorem 3 (F. RIESZ’ representation thm) Let l : Lp(A) → R be a continuous

linear functional, 1 ≤ p < ∞. Then there is a unique g ∈ Lq(A) such that

l( f ) =

A

f (x)g(x) dx

for all f ∈ Lp(A). Moreover, ‖l‖∗ = ‖g‖q,A.

Proof: The Radon-Nikodym theorem is used in most proofs of the represen-

tation theorem. A more direct proof in the one dimensional case is given in [R.

p.121].

Remark:

1) There are continuous linear functionals in L∞(A) that do not have this simple

form.

2) One says that Lq is the dual of Lp (1 ≤ p < ∞).

6

1.4 Weak convergence in Lp

Weak convergence in Lp is important for many applications, for example, it leads

to the existence theorems for partial differential equations. The so-called direct

methods in the Calculus of Variations are based on this concept.

Suppose that u1, u2, u3, . . . are functions in Lp. We say that uk → u weakly in

Lp, if

limk→∞

l(uk) = l(u)

for every continuous linear functional l : Lp → R . Using Riesz’ representation

theorem we can state the definition in a more convenient form.

Definition 4 Let 1 ≤ p < ∞. Suppose that u1, u2, . . . and u are functions in Lp(A).

We say that uk → u weakly in Lp(A), if

limk→∞

A

ukv dx =

A

uv dx for each v ∈ Lq(A).

Remark:

1) Sometimes the notation uk u is used to indicate weak convergence.

2) Strong convergence implies weak convergence: if ‖uk − u‖p → 0, then

uk u weakly in Lp.

3) The weak limit is unique, that is, unique in Lp.

Example: The functions un(x) = sin(nx) converge weakly in L2([0, 2π]) to zero.

By the Riemann-Lebesgue lemma

limn→∞

2π∫

0

v(x) sin(nx) dx = 0

for every v in L2([0, 2π]). (This follows easily from Bessel’s inequality.)

Example: (Warning!) Suppose that f : [a, b] → R is a bounded measurable

function, for example, assume 0 ≤ f (x) ≤ 2. Then there is a sequence of functions

vn : [a, b]→ 0, 2 converging weakly in L2([a, b]) to f . Observe that vn(x) = 0 or

= 2 ( vn takes no other values!!)

This is not difficult to realise in the special case f (x) ≡ 1.

7

Example: Define ui :]0, 1[→ R by ui = iχ]0,i−3[ for i = 1, 2, 3, . . . Then ui → 0

strongly in Lp([0, 1]), if 1 ≤ p < 3. We have ‖ui‖3+ε = iε /3 for every ε > 0. As

we shall see, weakly convergent sequences are bounded in Lp-norm. Therefore

u1, u2, . . . does not converge weakly in Lp(]0, 1[), if p > 3. —Show that ui 0

weakly in L3(]0, 1[)! (This convergence is not strong.)

If ui u weakly in Lp(A), then there is a constant M such that ‖ui‖p,A ≤ M ≤∞ for all i = 1, 2, 3, . . . This follows from the following lemma.

Lemma 5 Suppose that u1, u2, . . . are functions in Lp(A), 1 < p < ∞. If the

sequence ‖ui‖p,A is unbounded, there is a w ∈ Lq(A) such that

limk→∞

A

uik (x)w(x) dx = +∞

for some subsequence.

Proof: The function w is constructed and written down in [S, pp.25-28].

The most important fact about Lp-spaces seems to be the following weak com-

pactness property. (Not valid for p = 1).

Theorem 6 (Weak Compactness) Let u1, u2, . . . be functions in Lp(A), 1 < p <

∞. If there is a constant M such that ‖ui‖p,A ≤ M for each index i, then there exists

a function u ∈ Lp(A) such that uik u weakly in Lp(A) for some subsequence.

Moreover

‖ui‖p,A ≤ limk→∞‖uik‖p,A.

Proof: Advanced books on Functional Analysis usually contain a proof. For

example, the above mentioned book of Sobolev [S] gives a proof on pp. 29-30.

See also [A]. Usually, the lower semicontinuity comes as a by-product of the

proof, but the following simple argument also yields this property. Since |x|p is a

convex function of x, we have

|y|p ≥ |x|p + p|x|p−2x · (y − x), p ≥ 1.

Thus ∫

A

|uik |p dx ≥∫

A

|u|p dx + p

A

|u|p−1u · (uik − u) dx.

Now |u|p−1u is in Lq(A), and so the last integral approaches zero as k→ ∞ (by the

weak convergence!). This gives us the desired lower semicontinuity.

8

Remarks:

1) The theorem is not true for p = 1.

2) The existence of solutions to partial differential equations is often a direct

consequence of the weak compactness.

3) Let ‖ui‖p ≤ M for i = 1, 2, 3, . . .According to the BANACH-SAKS theorem

there are indices i1 < i2 < i3 < . . .and a function u in Lp such that

∥∥∥∥∥

ui1 + ui2 + . . . + uiν

ν− u

∥∥∥∥∥

p

−→ 0

as ν → +∞, that is, the arithmetic means converge strongly. (The Banach-

Saks theorem is valid also for p = 1.)

1.5 Approximation in Lp and some other things

The essential fact is that the functions in Lp can be approximated in the Lp-normby smooth functions, if 1 ≤ p < ∞. These are constructed as convolutions. Let us

first state some auxiliary results.

Lemma 7 (“Continuity in the Lp-norm”) Let 1 ≤ p < ∞. If f ∈ Lp(A), then

limh→0

A

| f (x + h) − f (x)|p dx = 0,

where f is regarded as 0 outside A.

The proof is not quite simple. Usually one uses the theorem of Lusin, valid for measurable

functions.

Theorem 8 (Lusin) Let f : A → [−∞,∞] be a measurable function that is finite a.e.

in A. Suppose that (A) < ∞. Given ε > 0, there is a compact set Kε ⊂ A such that the

restriction f |Kεis continuous and mes(A\Kε) < ε .

Proof: See, [EG, p.15].

If A is open, then Lusin’s theorem can be combined with the extension theorem of

Urysohn-Tietze [R, p.148].

9

Theorem 9 (Urysohn-Tietze) Suppose that A is open and mes (A) < ∞. Let K ⊂ A be a

compact set. If f : K → R is a continuous function, then there is a function ϕ ∈ C0(A)

such that ϕ(x) = f (x), when x ∈ K. Moreover, maxA|ϕ| = max

A| f |

Lemma 10 C0(Ω) is dense in Lp(Ω), 1 ≤ p < ∞. Here Ω is open and C0(Ω)

denotes all continuous functions with compact support on Ω. In other words, if

u ∈ Lp(Ω), then there are functions ϕk ∈ C0(Ω) such that

limk→∞‖u − ϕk‖p,Ω = 0.

Remarks:

1) Of course, this implies that C(Ω) is dense in Lp(Ω).

2) Functions in L∞(Ω) cannot, in general, be uniformly approximated by con-

tinuous functions.

3) If ϕ : Rn → R is continuous, then the closure in Rn of the set where ϕ(x) , 0

is called the support of ϕ. Thus

supp(ϕ) = x ∈ Rn |ϕ(x) , 0.

If supp(ϕ) is compact (it is closed by definition) and if supp(ϕ) ⊂ Ω, then

we say that ϕ ∈ C0(Ω). In this case the distance between supp(ϕ) and the

boundary ∂Ω is positive.

4) A deep result for Lp-functions is related to the Lebesgue points. Suppose that

f ∈ Lploc(R

n), that is, f ∈ Lp(B) for every ball B in Rn . Then

limr→o+

1

mes (B(x, r))

B(x,r)

| f (x) − f (y)|p dy = 0

for a.e. x.

Define

ρ(x) =

Ce− 1

1−|x|2 , |x| < 1

0, |x| ≥ 1

and choose the constant C > 0 such that∫

Rn

ρ(x) dx = 1. “Friedrichs’ mollifier” is

ρε(x) =1

εnρ

(x

ε

)

=

C

εne− ε2

ε2 −|x|2 , when |x| < ε,

0, when |x| ≥ ε .

10

The constant C depends only on the dimension n. Now we have

Rn

ρε(x) dx = 1, ε > 0. (1)

Observe that ρε ∈ C∞0 (Rn). The support of ρε is the closed ball |x| ≤ ε .The convolution

uε(x) = (ρε ∗ u)(x) =

Rn

ρε(x − y)u(y) dy

is well-defined for u ∈ L1loc

(Rn). If u is defined only in the domain Ω, then we

regard u as extended to zero outside Ω : u(x) = 0, when x ∈ Rn \Ω. Hence we can

calulate the convolution uε for any u in L1loc

(Ω).

Observe that uε is always a smooth function: uε ∈ C∞(Rn), if u ∈ L1loc

(Rn). For

differentiation we have the rule

Dαuε = (Dαρε) ∗ u. (2)

Here α = (α1, α2, . . . , αn) and

Dα=

∂α1+α2+...+αn

∂xα1

1∂x

α2

2. . . ∂x

αnn

.

If u was defined in the domain Ω, then the formula (2) holds for the original

(unextended) u at all points x with dist(x, ∂Ω) > ε . Analogously, if supp u ⊂ Ω,then uε ∈ C∞

0(Ω), when ε < dist(supp u, ∂Ω).

Lemma 11 If u ∈ Lp(Ω), 1 ≤ p < ∞, then uε ∈ Lp(Ω) and ‖ρε ∗ u‖p ≤ ‖u‖p.Moreover

limε→0+‖uε − u‖p = 0.

Lemma 12 If u ∈ C(Ω) and if K ⊂⊂ Ω is compact then

maxx∈K|uε(x) − u(x)| → 0 as ε→ 0+.

In other words, the convergence uε → u is uniform on compact subsets.

11

Proof: The locally uniform convergence for a continuous u follows form

u(x) − uε(x) = u(x) −∫

ρε(x − y)u(y) dy

=

|x−y|<ε

(u(x) − u(y))ρε(x − y) dy

|u(x) − uε(x)| ≤ maxy

|x−y|≦ε

|u(x) − u(y)| · 1 (ε < dist(x, ∂Ω))

in accordance with Weierstrass theorem ( a function that is continuous on a com-

pact set us uniformly continuous).

If u ∈ Lp(Ω), then

|uε| ≤∫

ρε(x − y)|u(y)| dy ≤∫

ρε(x − y)|u(y)|p dy

1p∫

ρε(x − y) dy

1q

︸ ︷︷ ︸

=1

and so∫

|uε(x)|p dx ≤∫ (∫

ρε(x − y)|u(y)|p dy

)

dx

=

|u(y)|p(∫

ρε(x − y) dx

)

︸ ︷︷ ︸

=1

dy =

|u(y)|p dy.

This proves the contraction ‖uε‖p ≤ ‖u‖p.For the Lp-convergence, we first note that if ϕ ∈ C0(Ω), then it is easily seen

that∫

|ϕ(x) − ϕε(x)|p dx ≤ sup|x−y|≦ε

|ϕ(x) − ϕ(y)|p (ε < dist(supp ϕ, ∂Ω))

and so ‖ϕ − ϕε‖p → 0. (Doing some more work, we could replace C0 by C ∩ Lp.)

Now

‖u − uε‖p ≤ ‖u − ϕ‖p + ‖ϕ − ϕε‖p + ‖=(ϕ−u)ε︷ ︸︸ ︷

ϕε − uε ‖p︸ ︷︷ ︸

≤‖ϕ−u‖p

and so

limε→0+‖u − uε‖p ≤ 2‖u − ϕ‖p.

12

Since C∞0

(Ω) is dense in Lp(Ω), we can choose ϕ ∈ C0(Ω) so that ‖u − ϕ‖p is as

small as we please. This concludes the proof for the Lp-convergence ‖u−uε‖p → 0.

As an application we mention the Variational Lemma.

Lemma 13 (Variational Lemma) Let u ∈ L1loc

(Ω), Ω denoting an open set in Rn .

If∫

Ω

u(x)ϕ(x) dx = 0

whenever ϕ ∈ C∞0

(Ω), then u = 0 a.e. in Ω.

Proof: Take x ∈ Ω and choose ε so small that 0 < ε < dist(x, ∂Ω). Then

ρε(x − x) will do as a test function so that

uε(x) =

u(x)ρε(x − x) dx = 0.

Let B be any closed ball in Ω, i.e. B ⊂⊂ Ω. Then

‖u‖1,B = ‖u − uε‖1,B → 0

as ε→ 0+. Hence u = 0 a.e. in B and hence a.e. also in Ω.

Remark: The lemma is fundamental in the Calculus of Variations. If u is con-

tinuous, the proof is more elementary and u ≡ 0 in this case. We shall need the

lemma to establish that the Sobolev derivatives of a function are unique up to sets

of measure zero.

Example: (Hermann WEYL) Let u ∈ L1loc

(Rn) and suppose that∫

u(x)∆ϕ(x) dx = 0

whenever ϕ ∈ C∞0 (Rn). Then u has a continuous representative which is a harmonic

function, i.e. the continuous u belongs to C2 and ∆u = ∂2u

∂x21

+ . . .+ ∂2u

∂x2n= 0. (In fact,

u ∈ C∞(Rn).)

Take ϕ(x) = ρε(x − z) for some fixed z. Then ∆uε(z) = 0. Indeed,

uε(z) = (ρε ∗ u)(z) =

ρε(z − x) u(x) dx

∆uε(z) = (∆ρε ∗ u)(z) =

∆zρε(z − x) u(x) dx

=

[∆xρε(z − x)]u(x) dx = 0

13

by assumption. Note that

∂2u

∂x2k

ρε(z − x) =∂2u

∂z2k

ρε(z − x)

by direct calculation (ρε(z − x) is a function of |z − x|2). Since z was arbitrary,

∆uε ≡ 0. Thus the function uε is harmonic.

One does not seem to get any further without using some deeper property of

harmonic functions. By the mean value property

uε(x) =1

mes (B(x, r))

B(x,r)

uε(y) dy (Actually uε = u!)

Since uε → u in L1(B(x, r)), we have

limε→0+

uε(x) =1

mes (B(x, r))

B(x,r)

u(y) dy.

For a.e. x, u(x) = lim uε(x), at least when ε approaches zero through a subse-

quence ε1, ε2, ε3, . . . (Corollary 2). Redefining u in a set of measure zero, we get a

function that satifies the mean value property. Hence (the redefined) u is harmonic

in Rn.

14

2 SOBOLEV SPACES

The situation with the derivatives belonging to some Lp-space was studied by

Tonelli, B. Levi, Sobolev, Kondrachev et consortes. The corresponding spaces are

named after Sobolev.

2.1 Wm,p and Hm,p

Throughout this chapter Ω denotes a domain or an open subset of Rn . Suppose

u ∈ C1(Ω), where Ω. Then integration by parts yields

Ω

u(x)∂ϕ

∂xk

dx = −∫

Ω

∂u

∂xk

ϕ(x) dx

when ϕ ∈ C∞0 (Ω). This formula is the starting point for the definition of weak

(distributional, generalized) partial derivatives.

Definition 14 Assume u ∈ L1loc

(Ω) . We say that vk ∈ L1loc

(Ω) is the weak partial

derivative of u with respect to xk in Ω if

Ω

u(x)∂ϕ

∂xk

dx = −∫

Ω

vk(x)ϕ(x) dx

for all ϕ ∈ C10(Ω). We write vk = Dku = ∂u

∂xkand ∇u =

(∂u∂x1, . . . , ∂u

∂xn

)

, provided the

weak derivatives exists.

The weak partial derivative is uniquely defined a.e. in Ω (by the Variational

lemma). Notice the requirement that the weak derivative is a function (to which

Lebesgue’s theory applies), not merely a distribution. It is sufficient to consider

all ϕ ∈ C∞0 (Ω) in the integration-by-parts formula.

We say that the function u belongs to the SOBOLEV SPACE W1,p(Ω), if u ∈Lp(Ω) and the weak partial derivatives

∂u

∂x1

,∂u

∂x2

, . . . ,∂u

∂xn

exist and belong to Lp(Ω) . Here 1 ≤ p ≤ ∞. We say that u ∈ W1,ploc (Ω) if

u ∈ W1,p(U) for each open U ⊂⊂ Ω . (In this case we may have ‖u‖p,Ω = ∞

15

or ‖∇u‖p,Ω = ∞.) For u ∈ W1,p(Ω) we define the norm

‖u‖W1,p(Ω) = ‖u‖1,p,Ω = ‖u‖p,Ω +n∑

k=1

∥∥∥∥∥

∂u

∂xk

∥∥∥∥∥

p,Ω

Any equivalent norm, as

‖u‖ =∫

Ω

[|u|p + |∇u|p] dx

1p

,

will do. We say that uk → u (strongly) in W1,p(Ω), if ‖uk −u‖W1,p(Ω) → 0. Provided

with this norm (or any equivalent norm) W1,p(Ω) is a Banach space.

Example: Let u(x) = |x| for −∞ < x < ∞. It is easily verified that

v(x) =

1, x > 0

−1, x < 0

is the weak derivative of u. Show that∫

ϕv dx = −∫

uϕ′ dx! (We can set v(0) = 0

or v(0) = A or even v(0) = ∞. Sets of measure zero do not count.)

Example: Let v(x) = 1, when x ≥ 0 and v(x) = −1, when x < 0. Then the weak

derivative of v does not exist in (−2, 2), for example. The origin is the crucial

point. (Dirac’s delta is not a function.)

Higher weak derivatives are defined in a similar way. If α = (α1, α2, . . . , αn), β =

(β1, β2, . . . , βn) are multi-indices, we write

|α| = α1 + α2 + . . . + αn, α! = α1!α2! . . . αn!(

α

β

)

=α!

(α − β)!β!=

n∏

k=1

(

αk

βk

)

xα = xα1

1xα2

2. . . xαn

n

Dα= D

α1

1. . .Dαn

n =∂α1+α2+...+αn

∂xα1

1∂x

α2

2. . . ∂x

αnn

In this notation

Dα(ϕψ) =∑

0≦|β|≦|α|

(

α

β

)

Dα−βϕ · Dβψ, LEIBNIZ’ RULE

ϕ(x + h) = ϕ(x) +∑

0<|α|Dαϕ(x) · hα

α!. TAYLOR’S FORMULA

16

If u ∈ L1loc

(Ω) and vα ∈ L1loc

(Ω) are related by

Ω

u(x) Dαϕ(x) dx = (−1)|α|∫

Ω

vα(x)ϕ(x) dx

for all ϕ ∈ C∞0 (Ω), we write vα = Dαu. This is a weak derivative of order |α|.We say that u ∈ Wm,p(Ω), if u ∈ Lp(Ω) and if Dαu exists and belongs to Lp(Ω)

for each multi-index α with |α| ≤ k. Provided with the norm

‖u‖Wm,p(Ω) = ‖u‖m,p,Ω =∑

|α|≤m

Ω

|Dαu|p dx

1p

,

the Sobolev space Wm,p(Ω) is a Banach space. The term with index α = (0, 0, . . . , 0)

is interpreted as

Ω

|u|p dx

1p

. The integer m counts the order of the highest weak

derivative.

Theorem 15 Assume that u ∈ Wm,p(Ω) for some 1 ≤ p < ∞. Then there exists a

sequence of functions ϕk ∈ C∞(Ω) such that ϕk → u in Wm,p(Ω), i.e.,

‖u − ϕk‖Wm,p(Ω)

k→∞→ 0.

About the proof. The case Ω = Rn is relatively simple. The general case was is in

[A, pp.52-53]. Let us just mention that the proof4 uses the partition of unity.

Partition of unity

The partition of unity is frequently used in the theory of distributions. Suppose that Ω ⊂∞∪j=1

U j where each U j is open. Then there are functions ϕ ∈ C∞0

(Rn) such that:

1) 0 ≤ ϕ ≤ 1.

2) ϕ1(x) + ϕ2(x) + . . . = 1 at each point x in Ω .

3) If K ⊂ Ω is any compact set, then only finitely many of the functions ϕk are not

identically zero in K.

4This was proved by N. Meyers and J. Serrin in 1964. However, this was not, as it were, the

first proof.

17

4) Each ϕ j ∈ C∞0

(Uk) for some k = k( j).

We define Hm,p(Ω) as the completion of C∞(Ω) in the norm ‖ ‖m,p,Ω. Often Hm

means Hm,2, p = 2 being the most important special case. To be more precise,

u ∈ Lp(Ω) belongs to the space Hm,p(Ω), if there are functions ϕi ∈ C∞(Ω) such

that ϕi → u in Lp(Ω) and Dαϕi is a Cauchy sequence in Lp(Ω) for each multi-index

α, |α| ≤ m.

It is not difficult to see that Dαu exists, |α| ≤ m, if u ∈ Hm,p(Ω) . Moreover,

Dαϕi → Dαϕ in Lp(Ω) . The central result is [A, pp.52-53]:

Theorem 16 Hm,p(Ω) =Wm,p(Ω), 1 ≤ p < ∞, m = 1, 2, 3, . . .

There is also a characterization of the Sobolev space in terms of integrated

difference quotients. To this end, let ei = (0, . . . , 1, . . . , 0) denote the unit vector

in the ith direction. If u ∈ W1,p(Ω), then

Ω′

∣∣∣∣∣

u(x + hei) − u(x)

h

∣∣∣∣∣

p

dx ≤∫

Ω

|Diu|p dx

for any subdomain Ω′ ⊂⊂ Ω, when 0 < h < dist(Ω′, ∂Ω).

(For smooth functions ϕ this follows from the identity

ϕ(x + hei) − ϕ(x)

h=

1

h

h∫

0

Diϕ(x1, . . . , xi + t, . . . , xn) dt

and the general case follows by approximation.)

Remark: For smooth functions in convex domains the formula

ϕ(y) − ϕ(x) =

1∫

0

[

d

dtϕ(x + t(y − x))

]

dt =

1∫

0

(y − x) · ∇ϕ(x + t(y − x)) dt

is the source of many useful inequalities.

Theorem 17 Let u ∈ Lp(Ω), 1 < p < ∞. Suppose that for any subdomain

Ω′ ⊂⊂ Ω we have

Ω′

∣∣∣∣∣

u(x + hei) − u(x)

h

∣∣∣∣∣

p

dx ≤ Kp < ∞

whenever 0 < h < dist(Ω′, ∂Ω). Then the weak derivative Diu exists and ‖Diu‖p,Ω ≤K.

18

Proof: See [GT, p.169].

In conclusion, there are three5 equivalent definitions for the Sobolev space:

I The definition based on the integration-by-parts formula.

II The definition based on approximation by smooth functions with respect to

the Sobolev norm.

III The characterization in terms of the integrability of difference quotients.

Often, II is used to prove auxiliary inequalities and imbedding theorems, III is

sometimes used to prove the existence of weak derivatives. But I is the Main

Definition.

When we say that u ∈ W1,p(Ω) is, for example, continuous, we mean that there

exists a continuous function ϕ such that u(x) = ϕ(x) for a.e. x ∈ Ω . Thus u can be

made continuous after a redefinition in a set of measure zero. (Remember that, in

general, functions in Lp are defined only almost everywhere.)

2.2 The Space W1,p

0(Ω)

We wish to introduce functions with boundary values zero in Sobolev’s sense.

Remember that functions in W1,p(Ω) can be approximated by functions in C∞(Ω)

with respect to the norm

‖u‖1,p,Ω = ‖u‖p,Ω +n∑

k=1

∥∥∥∥∥

∂u

∂xk

∥∥∥∥∥.

If the approximation can be done using merely functions with compact supportΩ,

then the function itself is in a closed subspace denoted by W1,p

0(Ω).

Definition 18 Suppose that u ∈ W1,p(Ω). We say that u ∈ W1,p

0(Ω), if, given ε > 0,

there is a function ϕε ∈ C∞0

(Ω) such that ‖u − ϕε‖1,p,Ω < ε .

Hence W1,p

0(Ω) is the closure of C∞

0(Ω) with respect to the corresponding

Sobolev norm. Clearly,

Cp

0(Ω) ⊂ W

1,p

0(Ω) ⊂ W1,p(Ω).

5IfΩ is the whole space Rn, a further definition is possible. It is based on the Fourier transform.

There is also a more advanced theory based on Bessel potentials.

19

There are functions in ⊂ W1,p

0(Ω), that do not have compact support in Ω . (Ex-

ample: n = 1, p = 2, Ω =]0, π[, u(x) = sin x, supp u = [0, π] ⊃ Ω .) If u ∈C(Ω) ∩W1,p(Ω) and if

limx→ξx∈Ω

u(x) = 0

at each boundary point ξ ∈ ∂Ω, then u ∈ W1,p

0(Ω). However, there are continu-

ous functions in the Sobolev space W1,p

0(Ω) that do not have “the right boundary

values” zero in the classical sense:

Example: Ω = x ∈ R3 | 0 < |x| < 1. Here |x| =√

x21+ x2

2+ x2

3. Then

u(x) = ln1

|x|, x ∈ Ω,

is in W1,2

0(Ω). The origin is an isolated boundary point. We have

limx→0

ln1

|x|= +∞ (, 0).

(At all the other boundary points the function has the right boundary values in the

classical sense.)

Remark: Let Ω ⊂ Rn and suppose that p > n. Then every function in W1,p

0(Ω) is

continuous and takes the boundary values zero in the classical sense. (In applica-

tions, one usually has p = 2 < 3 ≦ n, unfortunately.)

An extremely important property of the space W1,p

0(Ω) is that it is closed even

under weak convergence. That is, if u1, u2, . . . belong to W1,p

0(Ω) and if ui u and

∇ui ∇u weakly in Lp(Ω) (by definition u ∈ W1,p(Ω)), then u itself is in W1,p

0(Ω).

Let f ∈ W1,p(Ω) and suppose that u ∈ W1,p(Ω).We say that u has the boundary

values f in Sobolev’s sense, if u − f ∈ W1,p

0(Ω). (Sometimes this is written as

u ∈ f + W1,p

0(Ω).) —There is a theory of so-called Trace Spaces like Wk,α(∂Ω) for

general boundary value problems. In particular,the normal derivative ∂u∂n

makes sense.

Lemma 19 If u ∈ W1,p

0(Ω), then the function uχ

Ω+ 0χ

Rn \Ω is in W1,p

0(Rn).

If Ω ⊂ Rn, then the cases 1 ≤ p < n, p = n (the borderline case), and p > n

are very different in some essential estimates.

20

Theorem 20 (the Sobolev inequality) Suppose that u ∈ W1,p

0(Ω), where Ω is a

domain in Rn . If 1 ≤ p < n, then there is a constant C depending only on n and p

such that

Ω

|u|np

n−p dx

n−p

np

≤ C

Ω

|∇u|p dx

1p

. SOBOLEV

Thus u ∈ Lp∗(Ω); p∗ =np

n−p= the Sobolev conjugate.

Proof: Extending u as zero outside Ω we may assume that u ∈ W1,p

0(Rn).

By approximation it is sufficient to prove the inequality for ϕ ∈ C∞0

(Rn). Let

us, for instructive purposes, write down the proof in the two dimensional case

n = 2, x = (x1, x2). Multiply

ϕ(x1, x2) =

x1∫

−∞

∂ϕ(t, x2)

∂x1

dt, ϕ(x1, x2) =

x2∫

−∞

∂ϕ(x1, t)

∂x2

dt,

to get

|ϕ(x1, x2)|2 ≤∞∫

−∞

∣∣∣∣∣

∂ϕ(t1, x2)

∂x1

∣∣∣∣∣

dt1 ·∞∫

−∞

∣∣∣∣∣

∂ϕ(x1, t2)

∂x2

∣∣∣∣∣

dt2.

Integrate with respect to x1 :

∞∫

−∞

|ϕ(x1, x2)|2 dx1 ≤∞∫

−∞

∣∣∣∣∣

∂ϕ(t1, x2)

∂x1

∣∣∣∣∣

dt1

∞∫

−∞

∞∫

−∞

|D2ϕ| dx1 dx2.

Integrate with respect to x2 :

"

|ϕ|2 dx1 dx2 ≤"

|D2ϕ| dx1 dx2 ·"

|D2ϕ| dx1 dx2 ≤("

|∇ϕ| dx1 dx2

)2

.

Hence ‖ϕ‖2 ≤ ‖∇ϕ‖1. This proves the theorem in the case n = 2, p = 1.

For the general p, 1 < p < 2, set ψ = |ϕ|γ with γ > 1 as selected below. Then

the above inequality applied on ψ yields

"

|ϕ|2γ ≤ γ2

("

|ϕ|γ−1|∇ϕ|)2

HOLDER

≤ γ2

("

|ϕ|2γ)2

γ−12γ

("

|∇ϕ|2γγ+1

) γ+1γ

21

for γ > 1. Take 2γ =2p

2−p. Then, after some arithmetic,

("

|ϕ|2p

2−p

) 2−p

p

≤ γ2

("

|∇ϕ|p) 2

p

.

This is the desired inequality for 1 < p < n = 2.

See [EG, pp.138-140] or [GT, pp.155-156] for general n. A proof can also be

based on the formula

u(x) =1

ωn

Ω

(x − y) · ∇u(y)

|x − y|ndy

valid for u ∈ W1,10

(Ω).

Remark: The case n < p < ∞. Let u ∈ W1,p(Ω), p > n. For each cube Q ⊂ Ωwe have

|u(x) − u(y)| ≤2pn

p − n|x − y|1−

np ‖∇u‖p,Q

for a.e. x, y ∈ Q. Hence u is locally Holder continuous in Ω with Holder exponent

α = 1 − np. If Ω is bounded, then u ∈ C(Ω) and ‖u‖∞,Ω ≤ C|Ω |1/n−1/p‖∇u‖p,Ω.

The case p = n (borderline case6) is very special.

For functions in W1,p (but not in W1,p

0) the corresponding inequalities are more

involved than the Sobolev inequality and they do require some additional regular-

ity of the domain in question (balls, cubes, domains, smooth domains, Lipschitz

domains etc).

1) 1 < p < n, u ∈ W1,p(Q), Q = a cube in Rn . Then

‖u‖p∗,Q ≤ p∗n − 1

n‖∇u‖p,Q + |Q|−1/n‖u‖p,Q

(The gain is that p∗ =np

n−p> p.)

2) Poincare’s inequality for u ∈ W1,p(Ω), 1 ≤ p < ∞.∫

Ω

|u(x) − uΩ|p dx ≤ C(n, p)|Ω |p/n∫

Ω

|∇u(x)|p dx.

6The Trudinger-Moser inequality

Ω

exp

(

|u|c‖∇u‖n

)n/(n−1)

dx ≤ C mes(Ω)

holds.

22

Here Ω is a convex domain and uΩ =1|Ω |

Ω

u(x) dx is the average of u over

Ω .

3) Poincare-Sobolev: 1 ≤ p < n, u ∈ W1,p(Ω), Ω = convex.

Ω

|u(x) − uΩ|np

n−p dx

n−pnp

≤ Cn,p

Ω

|∇u(x)|p dx

1/p

These inequalities constitute the main bulk in the Sobolev Imbedding Theorem.

For example, Lq(Ω) ⊃ W1,p

0(Ω), if q = p∗, and, for a domain of finite Lebesgue

measure, this is valid if 1 ≤ q ≤ p∗. The Rellich-Kondrachev theorem is even

stronger.

Theorem 21 (Rellich7-Kondrachev) Let Ω be an arbitrary bounded domain in Rn

and consider the space W1,p

0(Ω).

1 ≤ p < n. Then W1,p

0(Ω) is compactly imbedded in Lq(Ω), where 1 ≤ q <

p∗ =np

n−p(q need not be conjugate to p). In practical terms, if ui ∈ W

1,p

0(Ω), and

‖ui‖1,p,Ω ≤ M, i = 1, 2, . . . ,

then there exists a function u ∈ W1,p

0(Ω) and a subsequence such that

‖u − ui‖q,Ω → 0 as i→ ∞

for each fixed q, 1 ≤ q < p∗.

p > n. Then W1,p

0(Ω) is compactly imbedded in Cα(Ω), α = 1 − n

p. Now

|u(x) − u(y)| ≤ K|x − y|α.

p = n. This is the borderline case.

The same is true for W1,p(Ω), if the boundary of Ω is sufficiently regular.

Remarks:

• The functions, not their derivatives, converge strongly in Lq(Ω).

• The convergence uiν u need not be strong in Lp∗(Ω), but uiν u weakly

also in Lp∗ (Ω).

7The case p = 2 is credited to Rellich.

23

2.3 About W1,p

The first order Sobolev space has some special properties not valid for higher derivatives.

If u ∈ W1,ploc (Ω), so does u+ and u−. As usually, u = u+ − u−, |u| = u+ + u−. Hence,

|u| ∈ W1,ploc

(Ω). We have

∇u+ =

∇u, when u > 0,

0, when u ≦ 0,

with similar rules for u− and |u|. For example, for |u| we use∫

(u2ν + ε

2)1/2∇ϕ dx = −∫

ϕuν∇uν

(u2ν + ε

2)1/2dx (ε , 0).

The passage to the limit under the integral sign can be justified [R, Ch.4,Thm.6]. Now

|ϕuν∇uν|(u1ν + ε

2)−1/2 ≤ |ϕ∇uν|.

Finally, letting ε→ 0, we have by the Dominated Convergence Theorem∫

|u|∇ϕ dx = −∫

ϕ[

1χu>0 − 1χu<0 + 0χu=0

]

∇u dx

i.e., ∇|u| = ∇u, when u > 0, = −∇u, when u < 0, and = 0, when u = 0.

Lemma 22 Let u ∈ W1,p(Ω). Then ∇u = 0 a.e. on any set where u is constant.

Lemma 23 u, v ∈ W1,p(Ω) =⇒ maxu, v, minu, v ∈ W1,p(Ω).

The last lemma is not true with W1,p replaced by Wk,p, if k ≥ 2. Example!?

3 The equation ∆u = 4αu3+ f (x), α ≥ 0

Let Ω be a bounded domain in Rn . Suppose that f ∈ L∞(Ω) and that α ≥ 0 is , for

simplicity, a constant. Fix a function ϕ ∈ W1,2(Ω). Usually, ϕ is continuous even

in Ω. It represents the boundary values.

Problem: Minimize the variational integral

I(u) =

Ω

[

1

2|∇u|2 + αu4

+ f (x)u

]

dx

among all u ∈ W1,2(Ω) with boundary values ϕ (in Sobolev’s sense, i.e., u − ϕ ∈W1,2

0(Ω)).

The problem is interesting even for the case ϕ ≡ 0. Before proving the exis-

tence of a unique solution, let us observe that

24

[I]

The function u ∈ W1,2(Ω), u − ϕ ∈ W1,2

0(Ω), is minimizing if and only if

∫[

∇u · ∇η + 4αu3η + f (x)η]

dx = 0 EULER-LAGRANGE EQN in weak form

for all test-functions η ∈ W1,2

0(Ω). (Hence a partial integration and the Variational

lemma leads to the Eqn

∆u = 4αu3+ f (x),

provided that u has a second derivatives.)

To see this, note that for any real number ε

I(u + ε η) = I(u) + ε

[∇u · ∇η + 4αu3η + f (x)η] dx +1

2ε2

|∇η|2 dx

+ ε2

α(6u2η2+ 4uη3 ε+η4 ε2) dx

(⇒) Suppose that u is minimizing. The u(x) + ε η(x) is admissible, and

I(u + ε η) ≥ I(u)

by assumption. We must have∫

(∇u · ∇η + 4αu3η + f (x)η) = 0. (If I(u) +

ε J + ε2(. . .) attains its minimum for ε = 0, then J = 0!!!)

(⇐) If Euler’s equation in weak form holds, then

I(u + 1η) = I(u) + 0 +1

2

|∇η|2 dx + α

η2[u2 · 6 + 4uη + η2] dx ≥ I(u),

since |∇η|2 ≥ 0 and 6u2+4uη+η2 ≥ 2u2 ≥ 0. This means that I(u+η) ≥ I(u),

in other words u is minimizing. (Any admissible function v can be written

as v = u + (v − u), v − u = η ∈ W1,2

0(Ω).)

[II]

The minimizing function u is unique (if it exists). Suppose that there are two

solutions, say u1 and u2. Then u1 − u2 and u2 − u1 are in W1,2

0(Ω). Choose the test

25

function η = u2 − u1 in the Euler equation for u2 and the test function u1 − u2 in

the Euler equation for u1. Adding the two equations we get (DO IT)

Ω

|∇u1 − ∇u2|2 dx + 4

Ω

α

≥(u1−u2)2 12

(u21+u2

2)

︷ ︸︸ ︷

(u31 − u3

2)(u1 − u2)︸ ︷︷ ︸

≥0

dx = 0.

Hence the first integral is zero and so ∇u1 = ∇u2 a.e. inΩ . Thus u1−u2 is constant

a.e. in Ω (you may wish to prove this in Sobolev’s space). The boundary values

will force this constant to be zero. This shows that u1 = u2.

[III]

There exists a minimizing function u. We need some estimates to begin with.

Now

∣∣∣∣∣

f (x)v dx

∣∣∣∣∣=

∣∣∣∣∣

f (x)(v − ϕ) dx +

f (x)ϕ dx

∣∣∣∣∣

HOLDER

≤ ‖ f ‖2‖v − ϕ‖2 + ‖ fϕ‖1SOBOLEV

≤ ‖ f ‖2CΩ‖∇(v − ϕ)‖2 + ‖ fϕ‖1≦ ‖ f ‖2CΩ‖∇v‖2 + ‖ f ‖2CΩ‖∇ϕ‖2 + ‖ fϕ‖1= A‖∇v‖2 + B

for all v. This means that

I(v) ≥1

2‖∇v‖2

[

‖∇v‖2 −A

2

]

− B

and so I(v) ≥ −A2

32− B. Hence we have

−∞ < Infv

I(v) ≤ I(ϕ) < +∞.

Observe also that, for example,

‖∇v‖2 ≤ max2 +A

2, B + I(v) ≈ I(v).

By the definition of the infimum, there are admissible functions u1, u2, u3, . . .

such that

limi→∞

I(ui) = I0 = inf I(v).

26

This is called a minimizing sequence. We may assume that

I0 ≤ I(ui) < I0 + 1.

By the above bound

‖∇ui‖2 ≤ max

2 + A2, 1 + B + I0

= K

for all i = 1, 2, 3, . . . Now

‖ui‖2 ≤ ‖ϕ‖2 + ‖ui − ϕ‖2 ≤ ‖ϕ‖2 +CΩ‖∇ui − ∇ϕ‖2 ≤ ‖ϕ‖2 + CΩK + CΩ‖∇ϕ‖2

for all i = 1, 2, 3, . . . By the weak compactness of L2(Ω) there are functions u and

w in L2(Ω) such that

uiν u, ∇uiν w

weakly in L2(Ω). We must have that w = ∇u and u ∈ W1,p(Ω). Since ui − ϕ ∈W

1,p

0(Ω), so does u − ϕ. Now we claim that I(u) = I0, that is, u is the solution.

It is sufficient to establish that

I(u) ≤ limv→∞

I(uν)

since I(u) ≥ I0 (u is admissible!). First,

Ω

|∇uiν |2 dx ≥∫

Ω

|∇u|2 dx + 2

Ω

∇u · (∇uiν − ∇u) dx

︸ ︷︷ ︸

→ 0by the weak

convergence in L2(Ω)

and so

limν→∞

Ω

|∇uiν |2 dx ≥∫

Ω

|∇u|2 dx.

By the Rellich-Kondrachev theorem ui − ϕ → u − ϕ strongly in L2(Ω) at least

for some subsequence. Hence, passing again to some subsequence, we have that

ui(x) → u(x) at a.e. point x ∈ Ω . Thus

Ω

αu4 dx ≤ lim

Ω

αu4i dx

27

by Fatous Lemma, at least for a subsequence. Finally,

f (x)uiν dx→∫

f (x)u dx

by weak convergence. Collecting results, we have the desired semicontinuity.

1) Is u continuous? What about u ∈ C2(Ω), that is, is u a classical solution?

This is REGULARITY THEORY (de Giorgi, Moser, Nash).

2) Are the boundary values attained in the classical sense; limx→ξx∈Ω

u(x) = ϕ(x), ξ ∈

∂Ω? (This is true only in “regular domains.”)

3) Stability? What do small changes of the data cause?

References

[A] R. Adams, Sobolev Spaces, Academic Press, New York 1975.

[EG] L. Evans & R. Gariepy, Measure Theory and Fine Properties of Functions,

CRC Press, Boca Raton 1992.

[G] E. Giusti, Direct Methods in the Calculus of Variations, World Scientific,

Singapore 2003.

[GT] D. Gilbarg & N. Trudinger, Elliptic Partial Differential Equations of Sec-

ond Order, 2nd Ed., Springer-Verlag, Berlin 1983.

[J] J. Jost, Partial Differential Equations, Springer, New York 2002. —Chapter

7.

[LL] E. Lieb & M. Loss, Analysis, American Mathematical Society, Providence

1977.

[R] H. Royden, Real Analysis, 2nd Ed., MacMillan, New York 1970.

[S] S. Sobolev, Applications of Functional Analysis in Mathematical Physics,

Translations of Mathematical Monographs, American Mathematical Society,

Providence 1963, pp.25-28.

28


Recommended