On the unit disk D, conjugate harmonic functions exists, so u(z) is harmonic if and only if u=Ref for some complex analytic function f(z).
We want to find similar reproducing formula for u from the analytic case.
Let ∑n=0+∞anzn be the power series expansion of f, then f1(θ)=f(eiθ)=∑n=0+∞aneinθ, this is the Fourier series of f1. Note that the Fourier coefficients of f is not symmetric because f is not real valued (otherwise has to be zero).
Similarly, we have the Fourer series expansion fr(θ)=∑n=0+∞anrneinθ. Here I write ′′=′′ because the series does converge uniformly.
By the convolution theorem, the Fourier coefficients of Cr should be rn, so Cr(θ)=∑n=0+∞rneinθ. One can directly verify this by the geometric series 1−z1=∑n=0+∞zn.
Now let u be a harmonic function on D. Then we can write u(z)=f(z)+f(z) for an analytic function f on D. Let f(z)=∑n=0+∞anzn be the power series of f. Then u(reiθ)=2Rea0+∑n=1+∞anrneinθ+∑n=1+∞anrne−inθ=∑n=−∞+∞cnr∣n∣einθ, where we take
Then we have ur(θ)=u1∗Pr(θ). From the convolution theorem we see that the Fourier coefficients of Pr(θ) should be r∣n∣, so Pr(θ)=∑n=−∞+∞r∣n∣einθ. The series converges absolutely (hence uniformly) so the equation is justified. We call it the Poisson kernel on the circle.
Lemma. Let φ be an integrable function on the circle, then φ∗Pr is a harmonic function on D.
Proof. Let H(z)=1−z1+z, then Pr(θ)=ReH(reiθ). Then uφ(reiθ) is the real part of g(reiθ)=2π1∫−ππφ(t)1−rei(θ−t)1+rei(θ−t)dt=2π1∫−ππφ(t)eit−zeit+zdt.
Then ∂zˉ∂g(z)=2π1∫−ππφ(t)∂zˉ∂eit−zeit+zdt=0 because eit−zeit+z is analytic on D.
To justify the switch of integral and derivative, use ∂zˉ∂=21(∂x∂+i∂y∂) and the following tip.
Alternative proof. The Poisson formula and convolution thoerem tells us that the Fourier series of uφ(z) is ∑n=−∞+∞cnr∣n∣einθ, where cn is the Fourier coefficients of φ. Since φ is integrable, ∣cn∣ is bounded by ∥φ∥L1. Then for r<1 the series converges uniformly. Moreover, if we take derivatives on r or θ term by term, the resulted series still converges uniformly, because its absolute value is controlled by series of the form ∑n∈Znqr∣n∣, which converges, this tells us we are allowed to take partial derivatives term by term. Each term in the series looks like either cnzn or c−nzn, which are analytic or anti-analytic, so harmonic. Then Δuφ(z)=∑n=1+∞Δ(cnzn)+∑n=−∞−1Δ(cnzn)+Δ(c0)=0.
From the lemma we know from a (for example, integrable) function φ on D, we can produce a harmonic function uφ on D. But it is not clear so far whether the uφ we get are compatible with φ on the boundary. In this section we study boundary behavior of harmonic functions on the disk.
The Poisson kernel is an approximate identity when r→1.
Proof.
To show 2π1∫−ππPr(θ)dθ=1, use its Fourier series expansion term by term integrate. You can change integral and summation because the series converges uniformly.
Note that Pr is positive. For δ>0, supθ∈[−π,π]∖(−δ,δ)Pr(θ)≤Cδ⋅(1−r2) for some constant Cδ depending on δ. Then Pr(θ) converges to 0 uniformly when r→1. Then ∫(−π,π)∖(δ,δ)Pr(θ)dθ→0 when r→1.
Theorem. Let φ be an integrable function on the unit cirlcle.
(1) uφ(reiθ)=2π1∫−ππφ(t)1−2rcos(θ−t)+r21−r2dt is a harmonic function on D, such that limr→1uφ(reiθ)=φ(θ) when θ is continuous point of φ.
(2) Let φ be a continuous function on the unit circle. Then uφ(reiθ)=2π1∫−ππφ(t)1−2rcos(θ−t)+r21−r2dt is the unique solution to the Dirichlet problem, and ur→φ uniformly when r→1.
Proof. Just need to prove uniqueness. Suppose u~ is another solution, let v=uφ−u~. Then v is harmonic in D and continuous on Dˉ (the closed unit disk), and v=0 on the unit circle. Now suppose uφ(z0)=u~(z0) for some z0∈D, then ∣v(z0)∣>0. By maximum principle, on every C(r) for ∣z0∣<r<1, there exists some points wr such that ∣v(wr)∣>∣v(z0)∣. Then there must be a subsequence of (wrn) such that wrn→w′∈C. Since v is continuous, v(w′)≥∣v(z0)∣>0. Contradicts to v being zero on C.
A vector field on R3 is a a vector valued function v(x)=(p(x),q(x),r(x)). There are two types of vector fields:
In the 2-dimensional case, a “surface” becomes a curve, the flux is given by the integral ∫CE⋅dn, where n is the outwards normal to the curve C. Let γ(t)=(x(t),y(t)) be a parametrization of the curve C on [0,1], then the surface integral is given by ∫01p(t)y′(t)−q(t)x′(t)dt. We also write it in the form ∫C−pdy+qdx.
On the other hand, if we want to calculate the change of electronic potentials along a curve C, then it is given by the line integral ∫CE⋅dr=∫01pdx+qdy.
Here is a breif summery of differential forms in 2-dimensions.
Let D be a domain with smooth boundary C=∂D. Then the Green theorem, says
∫∂Dpdx+qdy=∫D(∂x∂q−∂y∂p)dxdy.
Let ω be the one-form pdx+qdy, then dω=∂y∂pdy∧dx+∂x∂qdx∧dy=(∂x∂q−∂y∂p)dx∧dy. Then the Green theorem can be written as
∫∂Dω=∫Ddω for a one form ω.
Quite naturally, we can identify the gradient field E=∇U as the 1-form dU=∂x∂Udx+∂y∂Udy. Then when we calculate the change of potential along curve C, it is just given by ∫CdU. Since we have d(dU)=0 by direct calculation, when C=∂D is a closed curve, ∫CdU=∫Dd(dU)=0, this explains the nature that change of potential does not depend on path chosen.
On the other hand, when we try to calculate the electronic flux outside D, we do ∫C−pdy+qdx, so it seems we should identify E with the 1-form −pdy+qdx in this case. This motivates us the following operator
Let ω=pdx+qdy be a 1-form. Define the ∗-operator by ∗ω=pdy−qdx. Then we have ∗∗ω=−ω.
The Fourier’s law says the heat flux is proportional to the gradient of temperature. In terms of vector fields, this is simply ∇u=−κF, where F is the heat flux. But from the point of view of differential forms, if we denote ω by the heat flux so that the amout of heat flowing outside D is given by ∫∂Dω, then the Fourier law should be intepreted as ω=−κ∗du.
By first law of thermaldynamics, increasing of temperature on D=∫D∂t∂u(x,t)dm(x)=−c∫∂Dω. Plug in Fourier law we get ∫D∂t∂u(x,t)dt=κc∫Dd∗du=κc∫DΔudx∧dy. Then we get the heat equation.
Using differential forms we can give a simple calculation of Laplacian in polar coordiantes. The exterior differential is coordinate invariant, so we always have du=∂x∂udx+∂y∂udy=∂r∂udr+∂θ∂udθ. And we know that d∗du=Δu(x,y)dx∧dy=Δu(r,θ)⋅rdrdθ. So it sufficies to calculate ∗ in polar coordinates.
From the Cartesian coordinate case, we have ∗dx=dy, ∗dy=−dx, so ∗∗dx=∗(dy)=−dx, ∗∗dy=∗(−dx)=−∗dx=−dy, we conclude that
Lemma.∗∗ω=−ω.
Note that this is independent of coordinates because we can define the ∗ operator in a coordinate-free way.
Now we calculate d∗du in polar coordinates. du=∂r∂udr+∂θ∂udθ, the main question is what is ∗dr and what is ∗dθ.
Let’s examine the function f=ln(x2+y2)=lnr2. Then df=∂r∂fdr+∂θ∂dθ=r2dr+0. On the other hand, in Cartesian coordinate df=∂x∂fdx+∂y∂fdy=x2+y22xdx+x2+y22ydy. ∗df=−∂y∂fdx+∂x∂fdy=−r22rsinθd(rcosθ)+r22rcosθd(rcosθ)=2(cos2θ+sin2θ)dθ=2dθ.
We see from this example that it should hold that 2dθ=∗df=∗(r2dr), so