An Introduction to Analysis/Differential forms

From testwiki
Jump to navigation Jump to search

In particular, the chapter covers subharmonic functions.

Implicit function theorem

4 Theorem A linear operator T from a finite-diemnsional vector space 𝒳 into itself is injective if and only if it is surjective.
Proof: Let e1,...en be a basis for 𝒳. The following are equivalent: (i) T has zero kernel. (ii) 0=T(j=1najej)=j=1najT(ej) implies that all the aj are zero. (iii) T(e1),...T(en) is a basis for 𝒳. Since the range of T is the span of the set {T(e1),...T(en)}, the theorem now follows.

4 Theorem Let Ω be a neighborhood of a point (a,b)ℝn×ℝm. If fj(a,b)=0 and fjπ’ž1(Ω) for j=1...n, and if the matrix

[f1x1f1xnfnx1fnxn]

is invertible at (a,b), then the equations fj(x,y)=0,j=1...n, has a unique solution x such that x(b)=a and x is π’ž1 in some neighborhood of b.
Proof (from [1]):

We need

4 Lemma If a linear operator T is injective in Ω, then T1 is defined and continuously differentiable in T(Ω).

Let F(x,y)=(f(x,y),y) for (x,y)Ω.

Connected spaces

The space A at top is connected; the shaded space B at bottom is not.

A set

E

is connected if there exists no open cover of

E

consisting of two disjoint open sets.

A connected component of a set E in G is the "maximal" connected subsets containing E; that is, the component = connected set G containing E. Every topological space, in other words, consists of components, which are necessarily disjoint and closed. That a topological space consists of exactly one component is equivalent to that the space is connected.

To give an example, induce to an arbitrary set G a topology as a collection of any subsets of G (i.e., the finest topology). The topological space G has no closed sets since every open set in G is also closed. The components of G are the same as all the subsets of G since .

4.3 Theorem The following are equivalent. Given a topological space G,

  1. G is connected.
  2. If G=AB, then both AB and AB are nonempty.
  3. Only and G have empty boundary.

Proof: Suppose G=AB for some sets A and B. If A and B are disjoint, so are A and B since AA. This is to say that (1) is false, which also follows if A and B are disjoint for the same reasoning. This shows that (1) implies (2). Now suppose E is nonempty, open, closed subset of G that is not G. Then so is GE. Thus, G=E(GE), the disjoint union of an open set and a closed set. This contradicts (2). Hence, (2) implies (3. Finally, suppose (1) is false; that is, there are at least two components of G, either of which has empty boundary but is not G.

A path is a continuous function from [0, 1] to some space; e.g., a straight-line represented by f(t) = A path is a loop if f(0) = f(1). e.g., a unit circle represented by f(t)=et2πi.

Two points a and b are said to be jointed by a path f if f(0) = a and f(1) = b. We say the space is path-connected, the importance of which notion is the following.

5.1 Theorem A set E is path-connected set if and only if it is connected.

Two paths that are homotopic.

Two paths are said to be homotopic if FIXME.

We say a space is simply connected if every path in the space is homotopic to a point. For example, in the plane ℝ2, every circle centered at the origin is homotopic to the origin. But in ℝ2/0 the circle fails to be homotopic to the origin. Hence, the former is simply connected while the latter is not. We also see, in light of theorem 3.1, that every simply-connected space is connected.

5.1 Theorem Let E be a set. The following are equivalent.

  • (i) df=0 implies that f is constant for any fπ’ž1(E)
  • (ii) E is connected.

Partition of unities

4 Lemma (Urysohn) A topological space X is normal if and only if for any disjoint closed sets A and B there exists a continuous function f such that 0f1, f=0 on A and f=1 on B. Proof (from Urysohn's lemma):

4 Corollary A topological space X is completely regular if and only if there exists a continuous injection from X to a compact Hausdorff space with continuous inverse.

4 Theorem A Hausdorff space is paracompact if and only if it admits a partition of unity.

Sheaf theory

We say β„± is a pre-sheaf on a topological space M if for any open subsets UVW of M

  • β„±(U) is an abelian group, ring, module, etc.
  • If ρU,V:UV is a set-inclusion (i.e., ρ(x)=x), then:
β„±(ρU,V):β„±(V)β„±(U) is a group (resp. ring, module, etc.) homomorphism such that β„±(ρU,U) = idβ„±(U) and β„±(ρU,W)=β„±(ρU,V)β„±(ρV,W)

If in addition the following holds a pre-sheaf is called a sheaf: for open sets UjM and their union U

  • (a) If f,gβ„±(U) and f|Uj=g|Uj for all j, then f|U=g|U
  • (b) If fβ„±(Uj) and fj|UjUk=fk|UjUk for all j,k, then there exists a fβ„±(U) such that f|Uj=fj for all j.

Here, we used (and will use) the notation g|B=β„±(ρB,A)(g) for gβ„±(A) and BA. We shall also denote β„±(ρU,V) simply by ρU,V when no confusion is possible.

Given sheaves β„±,𝒒 a sheaf (resp. pre-sheaf) homomorphism ϕ:ℱ𝒒 is given by a group (or ring, module, etc) homomorphism ϕU:β„±(U)𝒒(U) such that ϕUβ„±(ρU,V)=𝒒(ρU,V)ϕV for any UV.

A sheaf homomorphism ϕ:ℱ𝒒 is called an isomorphism if there is another sheaf homomorphism ψ:𝒒ℱ such that ϕψ = identity = ψϕ. A word of caution: it is not necessary that if ϕ is surjective, then ϕU is surjective for every open subset U, though the converse holds.

4 Corollary If ϕ is a sheaf homomorphism, then its kernel ker(ϕ) is a sheaf when it is defined by

ker(ϕ)(U)=ker(ϕU) for each U.

Proof: Everything is clear except that fker(ϕ)(U=Uj) (while we know fβ„±(U)) in (b) of the definition. But since (ϕUf)|Uj=ϕU(f|Uj)=ϕUfj=0 for all j and 𝒒 is a sheaf, we have ϕUf=0.

Though im(ϕ) can be defined similarly, in general it need not be a sheaf. Thus, we shall use the following formal procedure.

4 Lemma (sheafification) Every pre-sheaf β„± on M gives rise to a sheaf 𝒒 on M defined for each open subset UM by 𝒒(U)= all the maps f:UzUβ„±z satisfying the following:

  • U is the union of open sets VjU, and
  • For every j there is some gjβ„±(Vj) such that fz=(gj)z for all zVj (where fz=f(z) and (gj)z=(Vj,gj)).

If ψU(g)=g for every gβ„±(U), ψU is an injection from β„±(U) to 𝒒(U), and it is an isomorphism if β„± is actually a sheaf.
Proof: It is clear that 𝒒 is a sheaf and that ϕU is injective since we can just take Vj=U. Next, suppose β„± is a sheaf, let f𝒒(U) be given, and let Vj and gj correspond to f as in the definition. For every j,k we have

(gj|VjVk)z=fz=(gk|VjVk)z for all zVjVk,

which implies gj|VjVk=gk|VjVk. Since β„± is a sheaf, we can thus find gβ„±(U) with g|Vj=gj for every j. This is to say, we may have supposed that Vj=U in the first place. Hence, ψU is surjective.

Let ϕ be a sheaf homomorphism. By the application of the lemma to the pre-sheaf given by β„±(U)=im(ϕU) we define the image sheaf im(ϕ).

Corollary For any U, {x;xU,f(x)=g(x)} is open for any f,gβ„±(U)</mah>.<br/>Asheaf<math>β„± on M is said to be flasque if ρU,M:β„±(M)β„±(U) is surjective.

4 Lemma Every short exact sequence of sheaves:

0β„±0β„±1β„±20

induces for every U an exact sequence

0β„±0(U)β„±1(U)β„±2(U)

where the map β„±1(U)β„±2(U) is surjective if β„±0 is flasque.
Proof: By δj we denote the map β„±jβ„±j+1. It is clear that ker(δ0(U))=0 if ker(δ0)=0. Let fker(δ1U)=ker(δ1)(U)=im(δ0)(U) be given. Since im(δ0)(U) is a sheafification, we can write U as the union of open subsets VjU so that: for each j there is a gjim(δ0U) such that fz=(gj)z for all zVj. That is, there are uj with δ0U(uj)=gj and we have:

δ0VjVk(uj|VjVk)=f|VjVk=δ0VjVk(uk|VjVk),

and uj|VjVk=uk|VjVk since δ0U is injective. Hence, there exists uβ„±0(U) with u|Vj=uj. It follows: (δ0Uu)|Vj=δ0Vjuj=f|Vj and so δ0Uu=f. It remains to show that δ1U is surjective when F0 is flasque. To this end, let fim(δ1)(U) be given, and as before let Vj,uj so that we have:

δ1VjVk((ujuk)|VjVk)=0.

Then since ker(δ1VjVk)=im(δ0VjVk) there are sjk such that:

(ujuk)|VjVk=δ0VjVksjk.

We shall construct vj with

(vkvj)|VjVk=δ0VjVksjk

to be merged

In this chapter, we shall prove (after some works are done) Cauchy's integral formula, first by the Stoke's theorem then again by the notion of the winding number.

6.1 Theorem There exists a partition of unity ϕi subordinate to the cover {Gj}; that is:

  • (a) ϕi is infinitely differentiable in every Gj.
  • (b) supp ϕj is in Gj.
  • (c) If x is in Gj, then 1Nϕj=1 for some N. (locally finite)

Proof: Let G = the union of all Gj. Choose gj in C(G) so that {all supp gj} covers G and 0gj1. (See the lemma for why this is possible.)
Let ϕ1=g1, ϕ2=(1g1)g2, ϕ3=(1g1)(1g2)g3 and so forth. If 1mϕj=11mgj for some m, then the computation gives: 1m+1ϕj=11m+1gj. Since ϕ1=1(1g1), by induction,

1ϕj=11gj, which is locally finite.

For x in Gj, some gj(x)=1. Thus, (c) holds and the others (a) and (b) are also true by construction.

We define the integral of a form θ over E by for a partition of unity ϕj subordinate to the locally finite cover {Ej} of E,

Eθ=Ejϕjθ.

6.1 Theorem If u is analytic in Ω, then:

f(z)=12πiΩf(ζ)dζζz   zΩ.

(See also: Calculus:Complex_analysis)

We say a function f satisfies the mean value property when:

f(z)=12π(z+ϵeiθ)dθ.

An analytic function is an archetypical example, for the property is the immediate consequence of Caucy's integral formula. If f has the mean value property, then, for one, f is harmonic, and for another, the maximal principle become applicable to it.

6.1 Theorem
If u is analytic in Ω, then the following are equivalent:

  • (a) u(k) (z) = 0 for all k.
  • (b) uω = 0 for some ωΩ open.
  • (c) u has a non-isolated zero.

and if any of the above is true, then

  • (d) uΩ = 0.

Proof: Let E={f:fω}. If u is in E, then its derivative:

uΛ™=limh0h1(f(x+h)f(x))

is 0 in Ω since ω consists of interior points, and so we may suppose x+h is ω. Thus, from (b), (a) follows. That (b) implies (c) is obvious since an interior point is non-isolated. To show (d), let Z be {zΩ:f(z)=0}. Then Z is closed since the inverse of f, which is continuous by the inverse theorem, maps a closed set {0} back to Z in Ω. Z is also open, which we can know by considering a power series expansion. Since Z is nonempty by assumption, (d) follows after (a). (FIXME: this is still a partial proof)

6 Theorem (Runge) Let Kβ„‚ be compact, and ω be an arbitrary open subset of β„‚ containing K. Then the following are equivalent:

(a) For any fπ’œ(K) and an integer j, we can find a uπ’œ(ω) so that:
supK|fu|<2j
(b) K is holomorphically convex.

Proof: The theorem is a consequence of the Hahn-Banach theorem.

A compact subset K of a complex plane is said to have the Runge property if K satisfies any of the statements in the theorem.

6.2 Theorem (Weierstrass) Let Ωβ„‚ be open. Let the sequence zjΩ be discrete, and nj be a sequence of arbitrary integers. Then there exists a nonzero fπ’œ(Ω{z1,z2,...}) such that for each j (zzj)(nj)f is nonzero and analytic in some open set containing zj.
Proof: Let Kj be an exhaustion by compact sets of Ω with the Runge property. By the Runge property, for each j, we find a ujπ’œ(Ω) so that:

supKj|(zzj)nj+uj|<2j

where since the sequence zj is discrete, we may suppose zk∉Kj for any kj. Let

g=1(zzj)nj+uj, and f(z)=e0zg(s)ds.

Then f is analytic in Ω except for all zj. Also, let j be fixed and ω be an open set containing zj and no other terms in the sequence. Then fΛ™f=g in ω. Thus, by Cauchy's integral formula,

2πinj=ωg(s)ds=ωfΛ™(s)f(s)ds

It now follows that the argument principle says f has a zero of order nj (if the order is negative, then it is actually a pole).

This formulation is probably more illustrative, if it states more weakly.

6.2 Corollary Every discrete subset of Ωβ„‚ is the zero and pole set of some analytic function.
Proof: Every discrete set is countable.

6 Theorem Let Ωℝn be open and connected and η be one-form. Then the following are equivalent:

(1) η is exact on Ω.
(2) γ=0 if γ is a closed path.
(3) abη is independent of path.

Proof: On Ω, if η is exact, then η=df for some zero-form f. It thus follow:

abη=γdf=f(γ(1))f(γ(0)).

If γ is a closed path, then γ(1)=γ(0) by definition, and hence, (2) is true. Let γ1 and γ2 be arbitrary paths from a to b. Then

γ1γ2η=0 if (2) is true.

Thus, (2) implies (3). Finally, show (3) implies (1). Let f(x)=0xη. Then df=1nfxidxi. For each i, if xx+hiη=g(x+hi)g(x),thensince<math>0x+hi0x=,

fxi(x) =limhi1hi(xx+hiη
=limhi1hi(g(x+hi)g(x)) =gxi(x)

Here the derivative of f does exist since the integral is independent of path. We conclude that df=1nfxidxi=η.

Unitary operators

A continuous linear operator U:𝒳𝒴 is said to be unitary if U is surjective and preserves norms; i.e., x𝒳=Ux𝒴 for all x.

4 Theorem Suppose 𝒳 and 𝒴 are Hilbert spaces. A continuous linear operator U is unitary if and only if U*U and UU* are identity on 𝒳 and 𝒴, respectively.
Proof: The direct part follows from the identity

U*Uxx𝒳2U*Ux𝒳22Ux𝒴2+x𝒳2

and the converse from the identity

Ux𝒴2=U*Ux,x𝒳=x𝒳2.

Stokes formula

4 Theorem (Stokes) If ωβ„‚ has boundary which consists of finitely many Jordan curves, then:

ωη=ωdη

Proof: (FIXME: To be written)

4 Corollary (Green) If ωβ„‚ has boundary which consists of finitely many Jordan curves, then we have:

ω(fΔggΔf)dxdy=ω(fxggxf)dy(fyggyf)dx.

Proof: d(fxggxf)dy=(f2x2gg2x2f)dxdy.

Harmonicity

Let Ωℝn. A function uπ’ž2(Ω) is said to be harmonic if

1n2uxj2=0 (the Laplace equation)

We also define the poisson kernel

P(x/R,y)=Cn1(1|x/R|2)|yx/R|n

where Cn is the volume of a unit ball in ℝn.

4. Theorem Let Ω=BallR. Then u is harmonic on and continuous on Ω if and only if

u(x)=P(x/R,y)u(Ry)dω(y).

Proof: Suppose u is harmonic on Ω. Then using the Green's function

u(x)=P(x/r,y)u(ry)dω(y) for |x|<r<R.

Letting rR gives the direct part. Conversely, if xΩ, then the second derivative of u = 0 since P is harmonic on Ω.

4. Corollary (mean value property) Let Ω=BallR and u be harmonic on and continuous on Ω. Then

u(0)=u(Ry)dω(y)Cn.

Proof: Let x=0 in the theorem. Then P(0,y)=Cn1.

4. Corollary (maximum principle) If Ωℝ and u:Ωℝ is harmonic on Ω and continuous on Ω, then for xΩ,

minbΩuu(x)maxbΩu

where if the equality holds at some xΩ, then f is constant in the component of x.
Proof: (i) Suppose Ω=BallR. Then for xΩ

u(x) =|y|=1P(x/R,y)u(Ry)dω(y)
supbΩu|y|=1P(x/R,y)dω(y)
=supbΩu

since

|y|=1P(x/R)dω(y)supbΩ=1 when |x||R|.

Likewise, u(x)supbΩu. Thus,

infbΩuu(x)supbΩu

where inf and sup are actually min and max, respectively since the continuity of u and the compactness of a closed ball. (ii) Suppose Ω is arbitrary. Let xΩ. From (i) it follows that u is constant on every open ball containing x. Since Ω is open, every component of Ω is open. Since an open set is the union of non-disjoint open balls, u is constant on the component of x.

4. Theorem Let u be continuous on Ωℝn. Then the following are equivalent:

  • (i) u is harmonic.
  • (ii) If δ0 is given,
    u(x+ry)dω(y)Cndμ(r)=u(x)dμ(r)
  • where d(x,bΩ)δ and supp(dμ)[0,δ].
  • (iii) If δ>0 is given, then (ii) holds.

Proof: The mean value property says:

u(x)=u(x+ry)dω(y)Cn

By integrating both sides we get:

u(x)dμ(r)=u(x+ry)dω(y)Cndμ(r)

Hence, (i) implies (ii). Clearly, (ii) implies (iii). Suppose (iii), and let B be an open ball with Bω. Let h be harmonic on B and continuous on B such that u=h on bB. If Ballδ,xΩ, then using (iii)

(hu)(x+ry)dω(y)Cndμ(r)=(hu)(x)dμ(r)

where hu=0 on the boundary of B. Since dμ(r) has non-zero measure, u=h on B. Thus, (iii) implies (i).

Cauchy's integral formula

4 Thorem Let G be a bounded open subset of β„‚ whose boundary is smooth enough that Stokes' formula is applicable. If uπ’ž1(G), we have:

u(w)=12πiGu(w)zwdz1πGuzΒ―(z)1zwdxdy for wG

4 Theorem Let μ be a complex-valued measure with compact support in β„‚ and define

u(w)=1zwdμ

Schwarz lemma

4 Lemma (Schwarz) If f is analytic and |f(z)|1 for all |z|<1 and f(0)=0, then we have:

|f(z)||z| for all |z|1

Moreover, if the equality in the above holds at some point w0, then f is proportional to z
Proof: The hypothesis means that we can write f(z)=zg(z). Furthermore, if 0<r<1, the maximum principle says

sup|z|r|g(z)|=sup|z|=r|g(z)|1r.

and g is constant if g=1 at some point on the circle |z|=r. Letting r1 completes the proof.

Addendum

A Lie algebra is an algebra whose multiplication, denoted by [,], satisfies

  • (i) [x,x]=0, and
  • (ii) [[x,y],z]+[[y,z],x]+[[z,x],y]=0

for all x,y,z. Under the assumption (ii) we see (i) is equivalent to

0=[x+y,x+y]=[x,y]+[y,x].

When given an algebra is associative; i.e., (xy)z=x(yz) we can turn the algebra into a Lie algebra by defining [x,y]=xyyx, called a commutator. Indeed, it is clear that [x,y] distributes over scalars and addition and the condition (i) holds. It then follows [[x,y],z]=[x,[y,z]][y,[x,z]].


Also, ex=f(un)

Template:Subject