Functional Analysis and Partial Differential Equations: Notes On
Functional Analysis and Partial Differential Equations: Notes On
These are short incomplete notes. They do not substitute textbooks. The
following textbooks are recommended.
2 Hilbert spaces 9
2.1 Definition and first properties . . . . . . . . . . . . . . . . . . 9
2.2 The Riesz representation theorem . . . . . . . . . . . . . . . . 13
2.3 Orthonormal systems . . . . . . . . . . . . . . . . . . . . . . . 15
3 Lebesgue spaces 21
3.1 Review of measure spaces . . . . . . . . . . . . . . . . . . . . 21
3.2 Construction of measures . . . . . . . . . . . . . . . . . . . . 23
3.3 Jensen’s and Hölder’s inequalities . . . . . . . . . . . . . . . . 24
3.4 Minkowski’s inequality . . . . . . . . . . . . . . . . . . . . . . 26
3.5 Hanner’s inquality . . . . . . . . . . . . . . . . . . . . . . . . 28
3.6 The Lebesgue spaces Lp (µ) . . . . . . . . . . . . . . . . . . . 29
3.7 Projections and the dual of Lp (µ) . . . . . . . . . . . . . . . . 30
3.8 Young’s inequality and Schur’s lemma . . . . . . . . . . . . . 32
3.9 Borel and Radon measures . . . . . . . . . . . . . . . . . . . . 37
3.10 Compact sets . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.11 The Riesz representation theorem for Cb (X) . . . . . . . . . . 43
3.12 Covering lemmas and Radon measures on Rd . . . . . . . . . 47
5 Linear Functionals 82
5.1 The Theorem of Hahn-Banach . . . . . . . . . . . . . . . . . 83
5.2 Consequences of the theorems of Hahn-Banach . . . . . . . . 86
5.3 Separation theorems . . . . . . . . . . . . . . . . . . . . . . . 88
5.4 Weak* topology and the theorem of Banach-Alaoglu . . . . . 89
5.5 The direct method in the calculus of variations . . . . . . . . 97
1 Introduction
Functional analysis is the study of normed complete vector spaces (called
Banach spaces) and linear operators between them. It is built on the struc-
ture of linear algebra and analysis. Functional analysis provides the natural
frame work for vast areas of mathematics including probability, partial dif-
ferential equations and numerical analysis. It expresses an important shift
of viewpoint: Functions are now points in a function space.
Let Ω ⊂ Rn be an open bounded set with smooth boundary. One of
the deepest results in Einführung in die PDG was that the Green’s function
g(x, y) provides a map
f →u
given by ˆ
u(x) = g(x, y)f (y)dy =: T f
so that
−∆u = f in Ω
u=0 on ∂Ω
whenever f is sufficiently regular. It is not hard to see that
T : L2 (Ω) → L2 (Ω)
1. Rn and Cn equipped with the Euclidean norm are real resp. complex
Banach spaces.
kT xkY ≤ 1 if kxkX ≤ δ,
δx
and hence, if x ∈ X, x 6= 0, then kxkX ≤ δ and
X
kxkX δx
kT xkY = T ≤ δ −1 kxkX
δ kxkX Y
Let
T x := lim Tn x.
n→∞
The convergence is uniform on bounded sets, and hence the limit T is con-
tinuous and in L(X, Y ). Moreover
Definition 1.6 (Dual space). Let X be a normed space. We define the dual
(Banach) space as X ∗ = L(X, K).
and
k(xj + yj )klp ≤ k(xj )klp + k(yj )klp .
provided we can devide by k(xj + yj )klp . There is nothing to show if this
quantity is 0 and it is finite whenever we sum over a finite number of indices.
Then a limit argument gives the full statement.
In particular we obtain the triangle inequality. One easily sees that
k(xj )klp = 0 implies (xj ) = 0 and
Thus the spaces lp are normed vector spaces. Now suppose that xn = (xn,j )
is a Cauchy sequence in lp . Then for every j, n → xn,j is a Cauchy sequence
in K. Let yj = limn→∞ xn,j and y = (yj ). Then, for every m > 1 (assuming
p < ∞, since l∞ = B(N))
N
xm kplp
X
ky − = lim |yj − xm,j |p
N →∞
j=1
N
X
= lim lim |xn,j − xm,j |p
N →∞ n→∞
j=1
Thus v → k.k is continuous with respect to |.|. It attains the infimum on the
Euclidean unit sphere (which is compact). This minimum has to be positive
and we call it λ−1 . Then
|v| = |v||v/|v|| ≤ λ|v|kv/|v|k = λkvk.
The two inequalities imply the equivalence of the norms k.k and |.| by choos-
ing √
C = max d max kek k, λ, 1 .
Thus every norm on Kd is equivalent to the Euclidean norm, and any
two norms are equivalent.
A Cauchy sequence vm = (vm,j ) with respect to k.k is also a Cauchy
sequence with respect to the Euclidean norm, hence it converges to a vector
v with respect to the Euclidean norm, and hence also kvm − vk → 0.
This proves the claim for Kd . Now let X be a vector spaces of dimension
d. Then there is a basis of d vectors, and a bijective linear map φ from Kd
to X. If k.kX is a norm on X then x → kφ(x)kX is a norm on Kd . Thus the
first part follows. Since φ(xn ) is a Cauchy sequence with respect to k.kX iff
(xn ) is a Cauchy sequence in Kd with respect to the second metric, and one
converges iff the second converges.This completes the proof.
[21.10.2016]
[26.10.2016]
Proof. Exercise
2 Hilbert spaces
2.1 Definition and first properties
Definition 2.1. Let X be a K vector space. A map h., .i : X × X → K is
called inner product if
Examples:
hx,yi
If y = 0 there is nothing to show, so we assume y 6= 0 and define λ = hy,yi .
Then
|hx, yi|2
0 ≤ hx, xi −
hy, yi
which implies the Cauchy-Schwarz inequality.
defines a norm.
Proof. Exercise
Then
1
hx, yi = kx + yk2 − kx − yk2 (2.8)
4
if K = R and
1
hx, yi = kx + yk2 − kx − yk2 + ikx + iyk2 − ikx − iyk2 (2.9)
4
otherwise defines an inner product such that the norm is the norm of the
preHilbert space. Vice versa: The norm of a prehilbert space defines the
parallelogram identity.
Proof. We begin with a real normed spaces whose norm satisfies the paral-
lelogram identity. We define
1
hx, yi = (kx + yk2 − kx − yk2 ).
4
Then
hx, yi = hy, xi.
Since by the parallelogram identity
hence
kx + y + zk2 =2kx + zk2 + 2kyk2 − kx − y + zk2
=2ky + zk2 + 2kxk2 − ky − x + zk2
and
1 1
kx+y +zk2 = kxk2 +kyk2 +kx+zk2 +ky +zk2 − kx−y +zk2 − ky −x+zk2
2 2
1 1
kx+y −zk2 = kxk2 +kyk2 +kx−zk2 +ky −zk2 − kx−y −zk2 − ky −x−zk2
2 2
We claim
hλx, yi = λhx, yi
for all x, y ∈ X and λ ∈ R. It obviously holds for λ = 1 by checking the
definition, and for all λ ∈ N be the previous step, hence for all λ ∈ Z. But
then it holds for all rational λ and by continuity for λ ∈ R.
We complete the proof for complex Hilbert spaces: We define
3
1X k
hx, yi = i kx + ik yk2
4
k=0
and observe that hix, yi = ihx, yi, hx, yi = hy, xi by definition, Rehx, yi is
the previous real inner product and Imhx, yi = Rehx, iyi.
Corollary 2.7. A normed space is a pre-Hilbert space if and only if all two
dimensional subspaces are pre-Hilbert spaces.
Proof. It is a pre-Hilbert space if and only if its norm satisfies the parallelo-
gram identity which holds if and only if the parallelogram identity holds for
all two dimensional subspaces.
kx − yk = d(C, K)
kx − yj k → d(C, K).
Then
kyn − ym k2 =k(x − yn ) − (x − ym )k2
=2kx − yn k2 + 2kx − ym k2 − k2x − (yn + ym )k2
≤2(kx − yn k2 + kx − ym k2 ) − 4d2 (C, K)
→0 as n, m → ∞
[26.10.2016]
[28.10.2016]
Rehx − y, z − yi ≤ 0 (2.10)
hx − p(x), zi = 0. (2.11)
y(t) = y + t(z − y) ∈ C
This implies (2.10) and also the converse. In the case that C is a closed
subspace (2.10) is equivalent to the orthogonality relation.
J(λx) = λJ(x).
y → Q(x, y) ∈ H ∗ .
hz(x), yi = Q(x, y)
Clearly z(x1 + x2 ) = z(x1 ) + z(x2 ) and z(λx) = λz(x) and we define the
continuous linear operator Ax = z. Since
RehAx, xi ≥ δkxk2
we obtain
kAxk ≥ δkxk.
It particular A is injective and the range is closed. If it is not surjective
there exists z with kzk = 1 and z is orthogonal to the range, i.e.
hAx, zi = 0
0 = hAz, zi ≥ δkzk2 .
Finally
dn 2 dn
(x − 1)n = 2n (x − 1)n = 2n n!.
dxn x=1 dxn x=1
and r r ˆ 1
2n + 1 2m + 1
Pn (x)Pm (x) dx = δn,m
2 2 −1
and the functions r
2n + 1
Pn (x)
2
are an orthonormal system.
[28.10.2016]
[02.11.2016]
Proof. We compute
ˆ 2π ˆ 2π
1 inx −imx 1
e e dx = ei(n−m)x dx = 0
2π 0 2π 0
if n 6= m and = 1 if n = m.
−u00n + qun = λn un
The functions un have n − 1 zeroes in (0, 1).If λ 6= λn for some j then (2.12)
has only trivial solutions.
Moreover
N
X
p(x) = λj xj
j=1
and
N
X
hx, xn i = hp(x), xn i = λm hxm , xn i = λn
m=1
and
∞
X
2
kp(x)k = |λn |2 .
n=1
Examples:
1. N is countable.
5. Q is countable.
6. QN is countable.
8. l2 (N) is separable.
Theorem 2.20. The following properties are equivalent for a Hilbert space
H which is not finite dimensional.
Thus
N
X
N → kx − hx, xn ixn k
n=1
This implies that the partial sums are a Cauchy sequence and we define
N
X
x := lim an xn
N →∞
j=n
Then
N
X
2
kxk = lim |an |2 = k(an )k2 .
N →∞
n=1
The map is clearly linear, surjective (by the previous part) and an isometry.
To complete the proof it suffices to prove that l2 (N) is separable. Al-
most by definition the (countable) union of the subspaces of dimension N of
sequence being 0 for indices > N are dense: The truncated series converge
in L2 . It suffices now to find a dense countable subset of RN . QN is an
obvious choice.
[02.11.2016]
[04.11.2016]
3 Lebesgue spaces
3.1 Review of measure spaces
Reference:
1. {} ∈ A
2. A ∈ A implies X\A ∈ A
∞
S
3. An ∈ A implies An ∈ A.
n=1
f −1 ((t, ∞]) ∈ A
Examples:
π s/2
φ(A) = 2−s rs
Γ( 2s + 1)
[04.11.2016]
[09.11.2016]
df + df − df +
(x) ≤ (y) ≤ (y)
dx dy dy
and for all z
df + df −
f (z) ≥ max{f (x) + (x)(z − x), f (x) + (x)(z − x)}.
dx dx
Proof. If x0 < x1 < x2 then
dF +
F (t) ≥ F (t0 ) + (t0 )(t − t0 ).
dt
Thus
dF +
µ({F ◦ f ≤ s}) ≤ µ({F (t0 ) + (t0 )(f − t0 ) ≤ s})
dt
+
and min{F ◦ f, 0} is integrable since x → F (t0 ) + dF dt (t0 )(f − t0 ) (which is
affine in f ) is integrable. Then
ˆ ˆ
dF +
F ◦ f dµ ≥ F (t0 ) + (t0 )(f − t0 )dµ
X X dt
ˆ
dF +
=F (t0 ) + (t0 ) f dµ − t0 = F (t0 ).
dt X
If 1 < p < ∞, then the equality implies that g = λ|f |p−2 f¯ almost everywhere
for some λ ∈ K with |λ| = 1.
Proof. We copy the proof basically from the one for the sequence space. As
there it suffices to consider f and g with kf kLp = 1 and kgkLq = 1 and prove
ˆ ˆ
1 p 1 q
|f ||g|dµ ≤ |f | + |g| dµ = 1.
p q
The inequality is strict unless
1 1
|f g(x)| = |f (x)|p + |g(x)|q
p q
almost everywhere, which implies |g| = |f |p−1 . Now
ˆ ˆ
f gdµ ≤ |f g|dµ ≤ kf kLp kgkLq
and in the case of equality all inequalities must be equalities. Hence |g| =
|f |p−1 . Now suppose for some integrable function h
ˆ ˆ
hdµ = |h|dµ.
and hence
h = λ|h|
almost everywhere. Back to our situation above this implies
g = λ|f |p−2 f¯
whenever f and g are p-integrable, with equality for p > 1 iff f and g are
linearly dependent.
Proof. We assume first that f is nonnegative and omit the absolute value.
We claim that
ˆ ˆ
p
y→ f (x, y)dµ(x) and H(x) := f (x, y)dν(y)
X Y
are measurable functions. This follows from the Theorem of Fubini if f resp
f p are µ × ν integrable, and by an approximation argument in the general
for almost all x and y. Since we must have also equality in the equality
above we must have
f (x, y) = α(x)β(y)
for some measurable function α and β.
For the last part we apply the first part with the counting measure on
Y = {0, 1}. The product measure is defined in the obvious fashion, even
without assuming that µ is σ finite. If f is p integrable then by the definition
of the integral
µ({x : |f (x)| > t}) ≤ t−p kf kpLp
Let
∞
[ 1
A= {x : |f (x)| + |g(x)| > }
j
j=1
Proof. We may assume that kgkLp ≤ kf kLp (otherwise we exchange the two)
and kf kLp = 1 (otherwise we multiply f and g by the inverse of the norm).
The first inequality follows from the following pointwise inequality: Let
We claim that
and by integration
We apply the inequality with r = kg|Lp and recall that kf kLp = 1. The left
hand side becomes
h i
(kf kLp + kgkLp )p−1 + (kf kLp − kgkLp )p−1 kf kLp
h i
+ (kf kLp + kgkLp )p−1 − (kf kLp − kgkLp )p−1 ) kgkLp
=(kf kLp + kgkLp )p + (kf kLp − kgkLp )p .
and the derivative vanishes only at r = R and changes sign there. Thus
and the reverse inequality if p > 2. This implies (3.3) for real f and g. We
claim that (3.3) holds for complex f and g. It suffices to consider f = a > 0
and g = beiθ . Since
[09.11.2016]
[11.11.2016]
holds. Moreover
kλf kLp = |λ|kf kLp
and
kf kLp = 0
if and only if f vanishes outside a set of zero
µ({f 6= 0}) = 0.
These functions are p integrable for all p. They are a subvector space.
Proof. The vector space property follows from the Minkowski inequality.
The other statements are obvious.
converges if F (x) < ∞. Let f be the limit if F (x) < ∞, and 0 otherwise. It
is measurable. Since max{f, fn } ≤ F we obtain by dominated convergence
ˆ
p
kf − fn kLp = |f − fn |p dµ → 0.
Let A = lim supn,m→∞ khn − hm kLp . This limsup is obtained along two
subsequences n, m → ∞. Let D = dist(f, K). Then
Then N (t) attains its minimum at t = 0 on the interval [0, 1]. We claim
that its derivative at t = 0 is
ˆ
d
N |t=0 = p Re |f (x) − g(x)|p−2 (f (x) − g(x))(ḡ(x) − h̄(x))dµ.
dt
This implies the assertion.
To calculate the derivative we assume that f, g ∈ Lp (µ) and define
N (t) = kf + tgkpLp .
Since ˆ
j(f )(|f |q−2 f¯) = |f |q dµ
is an isometric isomorphism.
and
ˆ p p
p p p
|g| p−1 |h| p−1 dµ ≤ k|g| p−1 k q(p−1) khk r(p−1) = kgkLp−1 p−1
q khkLr
L p L p
since
p 1 1 p 1
+ = (1 − ) = 1.
p−1 q r p−1 p
We denote Lp (Rd ) (or even Lp ) for Lp (md ) where md is the Lebesgue
measure.
is integrable and
ˆ
I(f, g, h) := f (−x)g(x − y)h(y)dm2d (x, y)
Rd ×Rd
satisfies
|I(f, g, h)| ≤ kf kLp kgkLq khkLr
and
I(f, g, h) = I(g, f, h) = I(f, h, g) = I(h, g, f ).
Proof. We assume 1 < p, q, r < ∞ since the limit cases are simpler, and
follow by obvious modifications. Measurability is a consequence of the the-
orem of Fubini. It suffices to prove the statement for nonnegative functions
since ˆ ˆ
f ghdm2d ≤ |f gh|dm2d
f (x − y)g(y)
is integrable and
´
f (x − y)g(y)dmd (y) if integrable
f ∗ g(x) :=
0 otherwise
0
defines a unique element in Lr (Rd ) and
for
1 1
+ 0 =1
r r
and all h ∈ Lr . Since then
1 1 1
+ + =2
p q r
and, by Fubini and Lemma 3.21
ˆ ˆ
f ∗ gh(x)dx ≤ |f | ∗ |g||h|dmd
ˆ
= |f (x − y)||g(y)||h(x)|dmd (x, y)
The map ˆ
p
L (ν) 3 f → T f := k(x, y)f (y)dν(y) ∈ Lp (µ)
where
1 1
+ =1
p q
if kf kLp = kgkLq = 1. If p = 1 or q = ∞ this is an immediate consequence
of the theorem of Fubini. So we assume 1 < p, q < ∞. It suffices to prove
(3.4) for nonnegative functions f , g and k where f and g are bounded and 0
outside a set of finite measure, since then (3.4) follows by an approximation
and monotone convergence.
For z ∈ C with 0 ≤ Re z ≤ 1, we define
|f |pz−1 f if f 6= 0
fz =
0 otherwise,
and ( p
(1−z)−1
|g| p−1 g if g 6= 0
gz =
0 otherwise .
Then for σ ∈ R,
kfiσ kL∞ = kg1+iσ kL∞ = 1,
Moreover
f 1 = f, g 1 = g.
p p
Notice that fz and gz are bounded and zero outside a set of finite mea-
sure. By dominated convergence
ˆ
z → H(z) = gz (x)k(x, y)fz (y)dµ × ν
[16.11.2016]
[18.11.2016]
Since uε (z) → 0 as | Im z| → ∞
The theorem of Fubini in the form stated holds for µ×ν with the smallest
σ algebra containing all cartesian product of measurable sets. The Lebesgue
measure restricted to the Borel sets is not complete. We can easily complete
σ algebras.
and
N
X
µ(K) ≤ µ(U ) ≤ µ(Uxj ).
j=1
µ(Uj ) → µ(K).
3. Since open sets are countable unions of closed sets every open set is in
F.
We define
G = {A : X\A, A ∈ F}
Then G contains complements of elements and countable unions of elements
of G. Hence it is a σ algebra containing all open sets, and thus it is the
Borel σ algebra. This implies the first claim.
It is open and
∞
[ [
B= (K̊j ∩ B) ⊂ K̊j \Cj = U
j=1
Moreover [
µ(U \B) = µ( (K̊j \Cj )\B) < ε.
Lemma 3.30. Let (X, d) be σ compact, and µ a Borel measure such that
any compact set is of finite measure. Then µ is Radon and it is outer regular.
Proof. Only inner regularity has to be proven since outer regularity fol-
lows then by Lemma 3.29. Let A be Borel with finite measure (why does
this suffice?). By Lemma 3.29 there exists a closedS set C ⊂ A such that
µ(A\C) < ε. Let Kj be compact subsets with X = Kj and Kj contained
in the interior of Kj+1 . Then
µ(C ∩ Kj ) → µ(C)
and C ∩ Kj is compact.
Let
Aj = {x : f (t) > tj }.
Then X
kf − (tj − tj−1 )χAj kL1 < ε
and it suffices to approximate a characteristic function of a measurable set
A of finite measure by a continuous function. Let ε > 0. By inner and outer
regularity there exists a compact set K and an open set U so that
K⊂A⊂U µ(U ) ≤ µ(K) + ε.
Then d(K, X\U ) := d0 > 0 and we define
fL (x) = max{1 − Ld(x, K), 0} ∈ C(X)
Then if d0 L ≥ 1
1
kfL − χA kLp < ε p .
If L is sufficiently large then supp f is compact. Thus continuous functions
with compact support are dense.
[18.11.2016]
[23.11.2016]
We define
g(x) = min{f (y) + 2Ld(x, y)}.
y
One easily checks that g has Lipschitz constant 2L, and the mimimum is
attained in Bδ (x) and
1. A is bounded.
Proof. By the proof of Lemma 3.33 the Lipschitz continuous functions are
dense. The countable union of separable sets is separable and its closure is
separable. Hence it suffices to prove that
Proof. Since Lipschitz continuous functions with compact support are dense
we argue as for Cb (X).
1. C is bounded.
2. For every ε > 0 there exists δ so that for all |h| < δ and all f ∈ C
for |h| < δ. This contradicts the previous inequality. Similarly we deduce
the third part.
Vice versa: Suppose that C ⊂ Lp (Rd ) is closed, bounded, and satisfies
the three claims. We choose a smooth
´ function η supported in the unit ball
with values between 0 and 1 and η = 1, define ηr (x) = r−d η(x/r) and we
fix ε > 0. Then there exists δ so that by Minkowski’s inequality and the
second assumption
[23.11.2016]
[25.11.2016]
3. As a consequence ˆ
|L(f )| ≤ |f |dµ
and we can extend L to L1 (µ), hence L ∈ (L1 (µ))∗ and there exists
σ ∈ L∞ (µ) so that ˆ
L(f ) = f σdµ
X N
X
L(f ) = L(fj ) ≤ µ∗ (Uj ).
j j=1
in X. We define
η̃j
ηj = .
ρ
Finally, if A, B are Borel sets with positive distance there exist disjoint
open sets V and W containing A resp. B. Then
µ(U ) = µ∗ (U )
for open sets. By construction µ is bounded on compact sets and thus its
restriction to Borel sets is a Radon measure.
This gives
λ(f1 ) + λ(f2 ) ≤ λ(f1 + f2 ).
Now let |g| ≤ f1 + f2 . We define
(
f1 g
f1 +f2 if f1 + f2 > 0
g1 =
0 otherwise
which gives
λ(f ) = λ(f1 ) + λ(f2 ).
We claim that ˆ
λ(f ) = f dµ.
[25.11.2016]
[30.11.2016]
so that
3
rn ≥ sup{r : Br (x) ∈ F, x ∈ An }.
4
We stop if An = { }. For simplicity we consider the case when the procedure
does not stop. Then whenever j ≥ n, we have rj ≤ 34 rn (otherwise we would
not have chosen Brn (xn )) and
rn + rj
|xj − xn | ≥ rn ≥
3
and the balls Brj /3 (xj ) are all disjoint and
[
A⊂ Bj .
I = {j : 1 ≤ j < k, Bj ∩ Bk 6= { }}.
Next we bound the number of large balls, i.e #(I\K). Let now i, j ∈
I\K, i 6= j. We will give a upper bound on
hxi − xk , xj − xk i
cos(∠(xk xi , xk xj )) = .
|xi − xk ||xj − xk |
4
3 ≤ ri < |xi − xj | < rj ≤ ri , ri < |xi | ≤ 1 + ri , rj < |xj | < 1 + rj .
3
xi
0 θ xj
Proof. We fix θ so that 1 − M1d < θ < 1 and claim that there is a finite
collection of disjoint balls Bj , 1 ≤ j ≤ M in F so that
M
[
µ (A ∩ U )\ Bj ≤ θµ(A ∩ U ).
j=1
and repeat the argument with F1 the subset of balls with center in A1 .
After the kth step the complement has a measure at most θj µ(U ∩ A). So it
remains to prove the claim. Let F0 be the subset of balls with radii at most
1. Then we apply the Besicovitch covering theorem and obtain Gm . Then
M
[d [
A∩U ⊂ B
j=1 B⊂Gj
and
Md
X [
µ(A ∩ U ) ≤ µ A∩U ∩ B .
j=1 B∈Gj
is continuous from the left and right if µ(Br (x)) > 0 by inner and outer
regularity. Thus also Dµ ν(x) and Dµ ν(x) are Borel measurable since we
can write them as inf ’s and sup’s over rational radii. Moreover by inner
and outer regularity we obtain the same Dµ ν(x) and Dµ ν(x) if we use closed
balls.
2. Dµ ν is Borel measurable.
[30.11.2016]
[02.12.2016]
and
ν(B ∩ {x : Dµ ν(x) > t}) ≥ tµ(B ∩ {x : Dµ ν(x) > t}).
By outer regularity (µ(B) = inf{µ(U ) : B ⊂ U }) and it suffices to prove the
assertion for B = U open. Let A = {x ∈ U : Dµ ν(x) < t}. Let
ν(A) ≥ tµ(A)
Then
tµ(R(s, t)) ≤ ν(R(s, t)) ≤ sµ(R(s, t))
which implies µ(R(s, t)) = 0. Since
[
{x : Dµ ν(x) < Dµ ν(x)} = R(s, t)
s<t,s,t∈Q
Proof. It suffices to consider the case µ(Rd ) < ∞ and ν(Rd ) < ∞. We have
seen that
µ({Dµ ν(x) = ∞}) = 0
and hence, since ν << µ, ν(({Dµ ν(x) = ∞}) = 0. In the same fashion
Then
∞
X ∞
X ˆ ∞ ˆ
m+1
ν(A) = ν(Am ) ≤ t µ(Am ) ≤ t µ({Dµ ν > s})ds = t Dµ νdµ
m=−∞ m=−∞ 0 A
and
∞
X ∞
X ˆ ∞ ˆ
m −1 −1
ν(A) = ν(Am ) ≥ t µ(Am ) ≥ t µ({Dµ ν > s})ds = t Dµ νdµ.
m=−∞ m=−∞ 0 A
We let now t → 1.
exists almost everywhere and we define f (x) = 0 if it does not exist. Then
f is in the equivalence class of f˜. If f˜ ∈ Lploc (µ) then f ∈ Lploc and
ˆ
−1
lim µ(Br (x)) |f (y) − f (x)|p dµ(y) = 0
r→0 Br (x)
almost everywhere.
and Dµ ν lies in the equivalence class. Now the first claim follows from
Theorem 3.48.
For every t, |f (x) − t|p is integrable. From the first part
ˆ
−1
lim µ(Br (x)) |f (y) − t|p dµ(y) = |f (x) − t|p
r→0 Br (x)
Proof. Let (X, d) be a complete metric space and Aj open dense sets. Let
x ∈ X and ε > 0. Let x1 ∈ A1 so that d(x, x1 ) < ε/3 and 0 < δ1 < ε/3 so
that B2δ1 (x1 ) ⊂ A1 . We pick recursively xn , δn so that d(xn−1 , xn ) < δn ,
2δn < ε/3n and B2δn (xn ) ∈ An ∩ Bδn−1 (xn−1 ).
By construction, d(xn−1 , xn ) < ε/(2 · 3n−1 ) and, if n < m, d(xn , xm ) <
ε P m−n −j ≤ 3ε and (x ) is a Cauchy sequence with limit y. Since
2·3n j=0 3 4·3n n
xm ∈ Bδn (xn ) for m ≥ n the same is true for y, and y ∈ An for all n.
Then
sup{kT kX→Y : T ∈ F} < ∞.
Proof. Let
Cn = {x ∈ X : sup kT xkY ≤ n}.
T ∈F
This set is closed since both map and normS are continuous, and Cn is an
intersection of closed sets. By assumption Cn = X. We claim that some
Cn has nonempty open interior. If not then the sets Un = S X\Cn are open
and dense, with nonempty intersection, a contradiction to Cn = X. Let
U ⊂ Cn0 be nonempty and open. It contains a ball Br (x0 ). If kxk < r then
[02.12.2016]
[09.12.2016]
Examples:
∂x H = δ0 .
3. Let
1/2 if |x| < −t
f=
0 if |x| ≥ −t
Then
ˆ 0 ˆ −t
2 2 1
(∂tt − ∂xx )f (φ) = (∂t2 − ∂x2 )φ(t, x)dxdt
2 −∞ t
=φ(0, 0) = δ0 (t, x),
Thus
−∆g = δ0 .
[09.12.2016]
[14.12.2016]
T (φ) = 0.
for x ∈ K. Then
N
X
Tφ = T (ηj φ) = 0
j=1
If ψ ∈ L1 (Rd ) then
φ ∗ Tψ (x) = φ ∗ ψ(x).
Moreover, if supp φ = K1 and supp T = K2 then
supp φ ∗ T ⊂ K1 + K2 = {x + y : x ∈ K1 and y ∈ K2 }.
Definition 4.19. Let T ∈ D0 (Rd ) and let S ∈ D0 (Rd ) with compact support.
We define their convolution by
(S ∗ T )(φ) = T (S̃ ∗ φ)
Example 4.20.
Let d > 2 and
2 Γ( d2 ) 2−d
g(x) = |x| ∈ L1loc (Rd ).
d − 2 π d/2
Then
−∆g = δ0
The same construction works for all differential operators with constant co-
efficients.
Lemma 4.21. Suppose that Tn → T in D0 (U ) and that K ⊂ U is compact.
Then there exists k and C so that
sup{|Tn (f ) − Tm (f )| : f ∈ XK , kf kC k ≤ 1} → 0
b
as n, m → ∞.
Proof. The proof of the first part is the same is for Lemma 4.8. So, given
K, there exist k ≥ 1 and C so that for any n
is compact in Cck−1 (K). Let ε > 0. Then there exist finitely many functions
fm ∈ Cbk (K) so that the ε balls in Cbk−1 centered at fm cover B̄1 (0). We
may assume that they are in Cc∞ by Lemma 4.6. There exists n0 so that
if n ≥ n0 . Then, for any f ∈ B̄1 (0) there exists fm such that kf −fm kC k−1 <
b
ε and hence for n ≥ n0 there hold
[14.12.2016]
[16.12.2016]
2. If g ∈ C ∞ and for any multiindex α there exist c|α| and κ|α| so that
|∂ α g| ≤ c|α| (1 + |x|)κ|α|
then gf ∈ S(Rd ).
then g ∗ f ∈ S(Rd ).
Proof. The first property follows from the definition. By the first property,
in order to prove the second property it suffices to show
kxα (g ∗ f )ksup .
We observe that
xj (g ∗ f ) = (xj g) ∗ f + g ∗ (xj f )
and the claim follows by induction on the length of α.
and
|Tn (φ) − T φ|
→ 0.
sup|α|+|β|≤k kxα ∂ β φksup
Proof. The proof is similar as that of Lemma 4.21. The existence of C and k
follows from the idea of the proof of Banach-Steinhaus theorem. The conver-
gence result follows from the compactness of the ball {φ ∈ S(Rd ) | sup|α|+|β|≤k kxα ∂ β φksup ≤
1} in the space {φ ∈ S(Rd ) | sup|α|+|β|≤k−1 kxα ∂ β φksup < +∞}, which is
easy to see if we notice that
sup kxα ∂ β φksup ≤ R−1 sup kxα ∂ β φksup((BR (0))C ) +Rk−1 kφkC k−1 (BR (0))
b
|α|+|β|≤k−1 |α|+|β|≤k
∂ α Tf = Tfα .
f kpLp (U )
X
α
kf kW k,p = k∂
|α|≤k
[16.12.2016]
[21.12.2016]
has norm 1.
Proof. Exercise.
ηj f → f in W k,p (U ).
which implies
d
X d
X
∂xi φk ∂xj (cof Dφ)kj = 0.
k=1 j=1
This implies the claim if det Dφ 6= 0. If det Dφ(x) = 0 we apply the reason-
ing to φ + εx and send ε to 0.
The second term vanishes by Lemma 4.39 and we continue, assuming that
φ is smooth (which requires an approximation argument)
d ˆ
X ∂xk
= (∂xk f ) det(Dφ−1 )dmd (x).
U ∂yj
k=1
Examples:
The Vandermonde matrix is invertible and we can solve this system. The
coefficients aj hence exist and depend only on k. Then
k+1 0
0
X
1− aj (−j)κ−κ −1 φ(κ ) (0) = 0.
j=1
[21.12.2016]
[23.12.2016]
Lp
f (. + tej ) − f (.) if p < ∞
→ ∂j f in
t D0 if p = ∞
f (tej + .) − f
k∂xj f kLp (Rd ) ≤ lim inf .
t→0 t Lp
for C 1 ∩ W 1,p functions. Density completes the argument for p < ∞. For
p = ∞ we use that by the previous argument
ˆ ˆ 1ˆ d
X
d
(f (x + h) − f (x))φ(x)dm (x) = ∂xj f (x + th)hj φ(x)dmd (x)dt
0 Rd j=1
and hence
f (x + h) − f (x)
k kL∞ ≤ k|∇f |kL∞ .
|h|
We define ˆ
fB = f dmd = (md (B))−1 f dmd .
B B
[23.12.2016]
[11.01.2017]
There has been an omission in the proof of Theorem 4.45: We have not
shown that the difference quotient converges in Lp to the derivative - only
as a distribution.
|f (x) − f (y)|
sup ≤ kf 0 kLp . (4.5)
1− p1
x6=y |x − y|
The Sobolev inequality and Morrey’s inequality are the versions of these
inequalities (4.4), (4.5) in higher space dimension.
Theorem 4.49 (Morrey). Let U be open. Suppose that p > d and f˜ ∈
W 1,p (U ). Every point is a Lebesgue point and the canonnical representative
f is continuous. There exists c depending on p and d so that the following
is true: Let x, y ∈ U with
|x − y| < dist(x, Rd \U ).
Then
1− dp
|f (x) − f (y)| ≤ c|x − y| k|∇f |kLp (B|x−y| (x)) .
Proof. The inequality follows from
1− pd
f (y) − fBR (x0 ) ≤ CR k|∇f |kLp (BR (x0 )) (4.6)
∞
X
|f (y) − fBR (x0 ) | = fB2−j R (xj ) − fB21−j R (xj−1 )
j=1
∞
X
≤ |fB2−j R (xj ) − fB21−j R (xj−1 ) |
j=1
∞
1− pd −j(1− pd )
X
≤cR 2 k|∇f |kLp (BR (x0 ))
j=1
Proof. Part I is an exercise and we prove Part II. Let f ∈ W 1,p . By Morrey’s
theorem 4.49 we know that every point is a Lebesgue point and there is a
uniformly continuous representative. By Theorem 3.51 there exists a set A
whose complement has zero measure so that every x ∈ A is a Lebesgue point
and ˆ
lim md (Br (x))−1 |∇f (y) − ∇f (x)|p dmd (y) = 0.
r→0 Br (x)
ˆ !1
p
1− dp
|v(y)| ≤cr |∇f (x + z) − ∇f (x)|p dmd (z)
Br (0)
ˆ !1
p
d −1 p d
=cr m (Br (x)) |∇f (z) − ∇f (x)| dm (z)
Br (x)
=o(r).
to arrive at
k
Y
k −k
kf k k ≤2 k∂j f kL1 (Rk ) .
L k−1 (Rk )
j=1
(d−1)p
Now let 1 < p < d. We apply the above inequality (4.7) to |f | d−p .
Then
(d−1)p (d−1)p
kf kLqd−p
(Rd )
=k|f | d−p k d
L d−1 (Rd )
ˆ
(d−1)p (d − 1)p d(p−1)
≤kD|f | d−p kL1 (Rd ) ≤ |f | d−p |Df |dmd
d−p
d(p−1)
(d − 1)p
≤ kf kLqd−p
(Rd )
kDf kLp (Rd )
d−p
[11.01.2017]
[13.01.2017]
where we put all the constants into c̃. If 1 ≤ p < d by Hölder’s, Poincaré’s
and Sobolev’s inequality, with 1q + d1 = p1 and
1 1
p − r
α= 1 1,
p − q
we have
1−α α
kukLr (U ) ≤ kukLp (U ) kukLq (U ) ≤ CkukW 1,p (U ) ,
F j , f ∈ C(U × R × Rd )
aij ∈ L∞ (U )
F j , f ∈ Lp (U )
for 1 ≤ i, j ≤ d. Then u ∈ W 1,p (U ) is a weak solution to
d
X d
X
∂i (aij (x)∂j u) = ∂j F j + f
i,j=1 j=1
if and only if
ˆ d
X d
X
ij
a (x)∂j u∂i φ − F j (x)∂j φ + f φ̄ dmd = 0
U i,j=1 j=1
u − g ∈ W01,p (U ).
Definition 4.54. We call the (aij )1≤i,j≤d elliptic if there exists κ > 0 so
that
Xd
Re aij (x)ξi ξj ≥ κ|ξ|2
i,j=1
u=g on ∂U
and it satisfies ˆ
Re A(u, u) ≥ κ |∇u|2 dmd
U
kukL2 (U ) ≤ ck∇ukL2
and hence
κk∇ukL2 ≤ kF kL2 + Ckf kL2 .
we complete the bound by a second application of the Poincaré inequality.
−∆u = 0 in U
u=g on ∂U
has a unique solution u ∈ C 1 (U ) ∩ C(U ) for every continuous g. We call
this solution classical.
A classical solution is not necessarily in W 1,2 , and a weak solution is not
necessarily in C(U ). In most important cases both solutions are identical.
[13.01.2017]
[18.01.2017]
5 Linear Functionals
In this section we will study the dual space X ∗ of Banach spaces X.
Examples:
y + rx0 , y ∈ Y, r ∈ R.
with the obvious lV being an upper bound for the chain. Now let (V, lV )
be a maximal element. If V = X we are done. Otherwise we obtain a
contradiction by the first step.
2. Re L(x) ≤ p(x).
Proof. We apply the real theorem of Hahn Banach to the real part, and
extend it to a complex linear map by Lemma 5.3. To complete the proof we
observe that L and L̃ are the same iff the real parts are the same.
2. kLkX ∗ = klkY ∗ .
Proof. We define
p(x) = klkY ∗ kxkX .
Then
Re l(y) ≤ p(y)
for all y ∈ Y . We apply the theorems of Hahn-Banach to obtain L ∈ X ∗ so
that L|Y = l and
Re L(x) ≤ p(x).
kLkX ∗ = klkY ∗ .
If x∗ ∈ X ∗ then
Proof. The first claim is a consequence of Lemma 5.6. The second statement
is an immediate consequence of the definition.
l(y + tx0 ) = t
Then
|l((xj ))| ≤ k(xj )kl∞
for every converging sequence. By Theorem 5.5 it has an extension L to l∞ .
Clearly L(ej ) = l(ej ) = 0. We claim that it cannot be represented in the
form
∞
X
L((xj )) = yj xj
j=1
for (yj ) ∈ l1 - if it were represented in this fashion then all yj would have to
vanish.
1 1
≤ Re x∗j (xj ) = Re(x∗ (xj ) + (x∗j − x∗ )(xj )) <
2 2
This is a contradiction.
W k,p (U ) 3 f → (∂ α f )|α|≤k ∈ Lp (U × Σk )
pK (x) ≤ CkxkX .
Lemma 5.15. Let X be a normed vector space and V convex, open with
/ V . Then there exists x∗ ∈ X ∗ with
0∈
Re x∗ (x) < 0 if x ∈ V
Proof. Exercise
1. { }, A ∈ τ .
is open.
2. The weak∗ topology is the weakest topology with this property. This
means that the open sets are the smallest subset of the power set, such
that all sets above are contained in it, and arbitrary unions and finite
intersections are contained in it.
and open sets are arbitrary unions of such sets. This follows from a multiple
application of the distributive law of union and intersection.
Examples.
Hence
Mj
N [
[
X=W ∪ Yjk
j=1 k=1
[20.01.2017]
[25.01.2017]
is the set of all ’maps’ which assign to each α the element of Xα . There are
the obvious projections
πα : X → Xα
Suppose that all spaces Xα are topological spaces. Let τ be the smallest
topology (subset of the power set) containing all preimages of open sets in
Xα under πα .
Lemma 5.22. The preimages of open sets in Xα under πα define a subbase
S of τ .
Proof. We define the collection of arbitrary unions of finite intersections of
such sets. Then arbitrary unions and finite intersections have this form.
Thus every open set in τ is a union of finite intersections of such sets. Thus
these sets are a subbase.
Proof. The statement follows for the weak* converging subsequences by the
uniform boundedness principle. We may consider weakly converging se-
quences as weak∗ converging sequences in X ∗∗ , hence they are bounded.
Thus
x∗n *∗ x∗ =⇒ kx∗ kX ∗ ≤ lim inf kx∗n kX ∗ (5.3)
n→∞
and
xn * x =⇒ kxkX ≤ lim inf kxn kX . (5.4)
n→∞
Proof 2. Let (xl ) be a dense sequence on the unit ball of X. Let (x∗j ) be a
bounded sequence of element of X ∗ . Then there is a subsequence such that
(x∗jk (x1 ))
for k ≥ l. Since x∗j is bounded the yl define a unique continuous linear map
x∗ : X → K. Moreover x∗j *∗ x∗ .
[25.01.2017]
[27.01.2017]
Proof. Both are the smallest topology such that x → x∗ (x) = J(x)(x∗ ) are
continuous for all x∗ ∈ X ∗ .
Thus J : Y → Y ∗∗ is surjective.
2. Let 1 < p < ∞, U ⊂ Rd , un , u ∈ Lp (U ). Then
To see this, assume that un * u. Then supn kun kLp < ∞ and for all
p
φ ∈ D(U ) ⊂ L p−1 (U )
ˆ ˆ
un φdx → uφdx.
Vice versa: Suppose that supn kun kLp < ∞. The same is true for every
subsequence. Then every subsequence (unj ) has a weakly converging
subsequence, unjk * ũ. Then for all φ ∈ D(U )
ˆ
(ũ − u)φdx = 0
um → 0 as distributions,
Then
ˆ ˆ ∞ ˆ ∞
|un |p dx = n e−p(nx) dx = e−px dx = ku1 kpLp .
0 0
Thus again un * 0 as n → ∞.
6. Let
− p1 −(x/n)
un = n e χ(0,∞) .
Again kun kLp = ku1 kLp ,
− p1
kun ksup = n →0
and as above un * 0 as n → ∞.
∆v = f
Lemma 5.34. Let U ⊂ Rd be open and bounded, 1 < p < ∞ and either
X = W01,p (U ) or X = W 1,p (U ) under the assumption that the Whitney
extension property holds. Let (un ) ∈ X be a bounded sequence. Then there
is a subsequence (un ) and u ∈ X satisfying
3. un converges to u in Lp (U ).
[27.01.2017]
[01.02.2017]
|x| + 1 |x| + 1 p
|v|(|x| + 1) ≤ F (x + v) ≤ c 1 + x + v ≤ 2p c(1 + |x|)p
|v| |v|
and we obtain the desired conclusion.
F : U × R × Rd → R ∪ {∞}
(u, P ) → F (x, u, P )
P → F (x, u, P )
P → F (x, u, P )
un → u in Lp (U )
Then
F (x, un (x), ∇u(x)) → F (x, u(x), ∇u(x))
almost everywhere. Moreover
C(1 + |∇u(x)|p )
P → F (x, un (x), P )
is convex. Hence
∂F ∂F
(x, un , ∇u(x)) − (x, u, ∇u(x)) p
→ 0.
∂Pj ∂Pj L p−1
Since
∇un − ∇u * 0 in Lp
we obtain
ˆ
∂F
lim (x, un , ∇u(x))∂j (un − u)dmd (x)
n→∞ ∂Pj
ˆ
∂F
= lim (x, u, ∇u(x))∂j (un − u)dmd (x) = 0.
n→∞ ∂Pj
F (x, u, P ) ≥ C|P |p .
u − g ∈ W01,p (U )
and
E(u) = inf{E(v) : v − g ∈ W01,p (U )}.
and hence
sup kuj kW 1,p (U ) < ∞.
j
Proof. We argue as for the Dirichlet integral, choose v ∈ D(U ) and define
η(t) = E(u+tv). Then η(t) ≥ η(0) = E(u). We claim that it is differentiable
with respect to t:
d
1 X ∂F ∂F
(F (x, u+tv, ∇u+t∇v)−F (x, u, ∇u)) → (x, u, ∇u)∂j v+ (x, u, ∇u)v
t ∂Pj ∂u
j=1
(u, P ) → F (x, u, P )
[01.02.2017]
[03.02.2016]
6 Linear Operators
6.1 Dual and adjoint operators
Definition 6.1. Let X and Y be Banach spaces, T ∈ L(X, Y ). We define
0
the dual operator T ∈ L(Y ∗ , X ∗ ) by
0
(T y ∗ )(x) = y ∗ (T x) for y ∗ ∈ Y ∗ , x ∈ X.
5. Consider K = R,
−∆u = f u|∂U = 0
T : L2 3 f → u ∈ L2
Then T 0 = T .
7. kT 0 kY ∗ →X ∗ = kT kX→Y
hT x, yi = hx, T ∗ yi
and
ker T = {x ∈ X : T x = 0}.
If X0 ⊂ X is a subspace then
ran T = (ker T 0 )⊥
Thus
ran T ⊂ (ker T 0 )⊥ .
Now suppose that
y ∈ (ker T 0 )⊥
By Hahn Banach there exists y ∗ such that y ∗ |ran T = 0 and
Tx = f
T 0 y ∗ = 0 =⇒ y ∗ (f ) = 0.
1. ran T is closed,
2. ran T 0 is closed,
3. ran T = (ker T 0 )⊥ ,
4. ran T 0 = (ker T )⊥ .
Proof. Step 1: ’1 ⇐⇒ 2’
Let ỹ ∗ ∈ (ran T )∗ . By Hahn-Banach we may extend it to y ∗ ∈ Y ∗ . Since
and
ky ∗ kY ∗ ≤ ckT 0 y ∗ kX ∗
and hence ran T 0 is closed.
Step 1b: Now suppose that ran T 0 is closed. We want to prove that ran T
is closed. Again it suffices to consider the case that ran T is dense in Y .
Suppose there is sequence yn → 0 such that yn ∈ / T (B1 (0)). By the
∗ ∗ ∗
separation theorems there exists yn ∈ Y with kyn kY ∗ = 1 such that
Thus
kT 0 yn∗ kX ∗ ≤ kyn kY kyn∗ kY ∗ = kyn kY → 0.
However, by the Open Mapping Theorem 4.3, there exists c > 0 such that
if x∗ ∈ ran T 0 there exists y ∗ ∈ Y ∗ so that
ky ∗ kY ∗ ≤ ckx∗ kX ∗ , T 0 y ∗ = x∗ .
Let y ∈ BrY (0). There exists x0 with kx0 kX < 1 and ky−T x0 kY ≤ r/2. Then
there exists x1 with kx1 kX < 21 and ky − T (x0 + x1 )kY < 2−2 r. Recursively
we obtain a sequence xj with kxj kX < 2−j and
N
X −1
ky − T xn kY < 2−N r.
n=0
Then
∞
X
y=T xn
n=0
and the range intersected with BrY (0) is closed. Hence the range is closed.
Step 2: ’1 ⇐⇒ 3’:
This is a consequence of Lemma 6.4.
Step 3: ’2 ⇐⇒ 4:
’2 ⇐= 4’ is trivial since (ker T )⊥ is closed.
y ∗ (y) = x∗ (x)
This is well defined since x∗ vanishes on the null space of T . Since we may
assume that T is surjective we obtain from the open mapping theorem that
for y ∈ Y there exists x ∈ X with T x = y and kxkX ≤ CkykY . Thus
[03.02.2016]
[08.02.2017]
2. The image of the closed unit ball is relatively compact (the closure is
compact).
Tj → T in L(X, Y )
Proof. Let S ∈ L(Y, Z), T ∈ L(X, Y ) and U ∈ (V, X) as above and (vj ) a
bounded sequence in V . Then (U vj ) is a bounded sequence in X since U is
bounded, (T U vjl ) is a converging subsequence in Y since T is compact and
(ST U xjl ) is a convergent sequence in Z since S is continuous.
Let Tj → T a convergent sequence of compact operators and let (xn ) be
a bounded sequence. For simplicity we assume that it is bounded by 1.
We claim that for every ε > 0 there exists a subsequence xnj so that
for l, m ≥ l0 . Thus
kT (xjl − xjm )kY < ε
for l, m ≥ l0 .
Now let T be an operator of finite rank. Let (xj ) be a bounded sequence.
Then (T xj ) is a bounded sequence in a finite dimensional space and there
exists a convergent subsequence.
An invertible linear operator is compact if and only if every bounded
sequence has a convergent subsequence, which by Heine-Borel holds iff the
space has finite dimension.
xj − T xj → x − T x ∈ ran(1 − T ).
dist(xnj , ker(1 − T )) ≥ nj .
1
Let x̃j = kxnj k xnj . Then
xnj − T xnj
x̃j − T x̃j = →0
kxnj k
T compact ⇐⇒ T0 compact
K = T B1 (0) ⊂ Y.
T 0 yjl ∈ X ∗
εkxk ≤ kx − T xk.
Lemma 6.14. Let X be a Banach space, T ∈ L(X) such that ker T = {0}
and T is not surjective. Then ran T n+1 ⊂ ran T n but ran T n+1 6= ran T n .
with T̃ compact. Then ker S n = {0} and the range of S n is closed by Lemma
6.11. If S is not surjective then by Lemma 6.14 there exists a sequence yn
so that kyn k = 1, yn ∈ ran S n and
1
dist(yn , ran S n+1 ) ≥ .
2
We claim that
1
kT yn − T ym k ≥
2
if n < m. This is a contradiction to compactness and it remains to prove
the claim. It follows from
T yn − T ym = yn − ym − S(yn − ym )
and
ym + S(yn − ym ) ∈ ran S n+1 .
Thus S is injective and surjective, and invertibility follows now from the
theorem of the inverse, Corollary 4.4.
x − Tx = f
[08.02.2017]
[10.02.2017]
such that every function in the range is supported in a fixed ball. By the
difference characterisation of Sobolev functions
1
sup sup kEf (. + h) − Ef kLp (Rd ) ≤ ckf kW 1,p (U )
kf kW 1,p (U ) ≤1 h6=0 |h| 0
and
sup kEf kLp (Rd ) ≤ ckf kW 1,p (U ) .
0
kf kW 1,p (U ) ≤1
By Theorem 3.39 and Corollary 4.46 the unit ball in W 1,p (U ) is compact in
Lp (U ).
−∆u = f
T : L2 (U ) → L2 (U ).
If z ∈
/ (0, ∞) the equation
− ∆u = zu + f (6.1)
−zu − ∆u = 0
which is equivalent to
ˆ X
d
0= ∂xj u∂xj v − zuv̄ dmd
j=1
This equation has only the trivial solution u = 0 unless z ∈ (0, ∞). Thus
(6.2) is uniquely solvable and the solution is the unique weak solution to
(6.1).
Then
T : L2 ((0, 1)) → W 1,2 ((0, 1))
and it is compact as an operator to L2 ((0, 1)) since the embedding is com-
pact. We claim that there is no eigenvalue. If
ˆ x
zf (x) = f (y)dy
0
then, if z 6= 0
f 0 = z −1 f, f (0) = 0.
By Gronwall’s lemma
−1 x
|f (x)| ≤ e|z| |f (0)| = 0.
(T − S)A = 1Y , A(T − S) = 1X .
Thus the operator T −S is invertible and hence the set of invertible operators
is open.
(T − λ)2k0 y = 0 = (T − λ)k0 y = x.
This we can apply to problem (6.1) and conclude that there is a monotone
sequence of eigenvalues, with ∞ as only possible limit, and (6.1) is uniquely
solvable unless z is an eigenvalue.