Lecture 4

Let D\mathbb D denote the unit disk, CC denote the unit circle, C(r)C(r) denote the circle of radius rr, let D(r)\mathbb{D}(r) denote the disk of radius rr.

In this lecture we study two questions:

  • Find representation of uu on D\mathbb{D} by φ.
  • Can we reproduce φ(reiθ)\varphi(re^{i\theta}) by limr1u(reiθ))\lim_{r\to 1}u(re^{i\theta}))?

Some complex analysis

The Cauchy integral formula can be rewritten as follows:

Let z=reitz = re^{it}, then f(z)=f(reit)f(z) = f(re^{it}). Let fr(θ)=f(eiθ)f_r(\theta) = f(e^{i\theta}), we obtain a function frf_r on the circle, then

fr(θ)=12π02πf1(t)eiteitreiθdt=12π02πf1(t)11rei(θt)dt=f1Cr(θ), f_r(\theta) = \frac{1}{2\pi}\int_{0}^{2\pi}f_1(t)\frac{e^{it}}{e^{it}-re^{i\theta}}dt = \frac{1}{2\pi}\int_{0}^{2\pi}f_1(t)\frac{1}{1-re^{i(\theta - t)}}dt = f_1*C_r(\theta),

where Cr(t)=11reitC_r(t)=\frac{1}{1-re^{it}}, called the Cauchy kernel.

The Poisson kernel

On the unit disk D\mathbb{D}, conjugate harmonic functions exists, so u(z)u(z) is harmonic if and only if u=Refu = Ref for some complex analytic function f(z)f(z). We want to find similar reproducing formula for uu from the analytic case.

Let n=0+anzn\sum_{n=0}^{+\infty} a_nz^n be the power series expansion of ff, then f1(θ)=f(eiθ)=n=0+aneinθf_1(\theta) = f(e^{i\theta}) = \sum_{n=0}^{+\infty}a_n e^{in\theta}, this is the Fourier series of f1f_1. Note that the Fourier coefficients of ff is not symmetric because ff is not real valued (otherwise has to be zero).

Similarly, we have the Fourer series expansion fr(θ)=n=0+anrneinθf_r(\theta)=\sum_{n=0}^{+\infty} a_n r^ne^{in\theta}. Here I write =''='' because the series does converge uniformly.

By the convolution theorem, the Fourier coefficients of CrC_r should be rnr^n, so Cr(θ)=n=0+rneinθC_r(\theta) = \sum_{n=0}^{+\infty}r^ne^{in\theta}. One can directly verify this by the geometric series 11z=n=0+zn\frac{1}{1-z} = \sum_{n=0}^{+\infty} z^n.

Now let uu be a harmonic function on D\mathbb{D}. Then we can write u(z)=f(z)+f(z)u(z) = f(z)+\overline{f(z)} for an analytic function ff on D\mathbb{D}. Let f(z)=n=0+anznf(z) = \sum_{n=0}^{+\infty} a_n z^n be the power series of ff. Then u(reiθ)=2Rea0+n=1+anrneinθ+n=1+anrneinθ=n=+cnrneinθu(re^{i\theta}) = 2Re a_0 + \sum_{n = 1}^{+\infty} a_n r^n e^{in\theta} + \sum_{n = 1}^{+\infty} \overline{a_n}r^n e^{-in\theta} = \sum_{n = -\infty}^{+\infty} c_n r^{|n|}e^{in\theta}, where we take

cn={2Rea0,n=0an,n>0an,n<0.c_n = \begin{cases} 2Re a_0, n = 0\\ a_n, n > 0 \\ \overline{a_{-n}}, n < 0\end{cases}.

Then we have ur(θ)=u1Pr(θ)u_r(\theta) = u_1 * P_r(\theta). From the convolution theorem we see that the Fourier coefficients of Pr(θ)P_r(\theta) should be rnr^{|n|}, so Pr(θ)=n=+rneinθP_r(\theta) = \sum_{n=-\infty}^{+\infty}r^{|n|}e^{in\theta}. The series converges absolutely (hence uniformly) so the equation is justified. We call it the Poisson kernel on the circle.

Expression of Poisson kernel

Next we find a simple formula for the Poisson kernel. We can rewrite the Poisson kernel as

Pr(θ)=n=1+rneinθ+n=1+rneinθ+1=(Cr(θ)1)+Cr(θ)1+1=Re(2Cr(θ)1))=Re(1+reiθ1reiθ)=1r212rcosθ+r2.\begin{split} P_r(\theta) &= \sum_{n = 1}^{+\infty} r^ne^{in\theta} + \sum_{n = 1}^{+\infty}r^{n}e^{-in\theta} + 1\\ &= (C_r(\theta) - 1)+\overline{C_r(\theta)-1} + 1 \\ &= Re(2C_r(\theta) - 1)) \\ &= Re(\frac{1+re^{i\theta}}{1-re^{i\theta}}) \\ &= \frac{1-r^2}{1-2r\cos \theta + r^2}. \end{split}

Lemma. Let φ be an integrable function on the circle, then φPr\varphi*P_r is a harmonic function on D\mathbb{D}.

Proof. Let H(z)=1+z1zH(z) = \frac{1+z}{1-z}, then Pr(θ)=ReH(reiθ)P_r(\theta) = ReH(re^{i\theta}). Then uφ(reiθ)u_{\varphi}(re^{i\theta}) is the real part of g(reiθ)=12πππφ(t)1+rei(θt)1rei(θt)dt=12πππφ(t)eit+zeitzdtg(re^{i\theta}) = \frac{1}{2\pi}\int_{-\pi}^\pi \varphi(t)\frac{1+re^{i(\theta - t)}}{1-re^{i(\theta - t)}}dt=\frac{1}{2\pi}\int_{-\pi}^\pi \varphi(t)\frac{e^{it}+z}{e^{it}-z}dt. Then zˉg(z)=12πππφ(t)zˉeit+zeitzdt=0\frac{\partial}{\partial \bar{z}}g(z) = \frac{1}{2\pi}\int_{-\pi}^\pi \varphi(t)\frac{\partial}{\partial \bar{z}}\frac{e^{it}+z}{e^{it} -z}dt = 0 because eit+zeitz\frac{e^{it}+z}{e^{it} - z} is analytic on D\mathbb{D}. To justify the switch of integral and derivative, use zˉ=12(x+iy)\frac{\partial}{\partial \bar{z}} = \frac{1}{2}(\frac{\partial}{\partial x} + i\frac{\partial}{\partial y}) and the following tip.

Alternative proof. The Poisson formula and convolution thoerem tells us that the Fourier series of uφ(z)u_\varphi(z) is n=+cnrneinθ\sum_{n = -\infty}^{+\infty} c_n r^{|n|}e^{in\theta}, where cnc_n is the Fourier coefficients of φ. Since φ is integrable, cn|c_n| is bounded by φL1\|\varphi\|_{L^1}. Then for r<1r<1 the series converges uniformly. Moreover, if we take derivatives on rr or θ term by term, the resulted series still converges uniformly, because its absolute value is controlled by series of the form nZnqrn\sum_{n\in \mathbb{Z}}n^{q}r^{|n|}, which converges, this tells us we are allowed to take partial derivatives term by term. Each term in the series looks like either cnznc_n z^n or cnznc_{-n}\overline{z^n}, which are analytic or anti-analytic, so harmonic. Then Δuφ(z)=n=1+Δ(cnzn)+n=1Δ(cnzn)+Δ(c0)=0\Delta u_{\varphi}(z) = \sum_{n = 1}^{+\infty}\Delta (c_nz^{n}) + \sum_{n = -\infty}^{-1}\Delta(c_n\overline{z^n}) + \Delta(c_0) = 0.

Boundary value of harmonic functions

From the lemma we know from a (for example, integrable) function φ on D\mathbb{D}, we can produce a harmonic function uφu_\varphi on D\mathbb{D}. But it is not clear so far whether the uφu_\varphi we get are compatible with φ on the boundary. In this section we study boundary behavior of harmonic functions on the disk.

Theorem

The Poisson kernel is an approximate identity when r1r\to 1.

Proof.

  • To show 12πππPr(θ)dθ=1\frac{1}{2\pi}\int_{-\pi}^{\pi}P_r(\theta)d\theta = 1, use its Fourier series expansion term by term integrate. You can change integral and summation because the series converges uniformly.

  • Note that PrP_r is positive. For δ>0\delta > 0, supθ[π,π](δ,δ)Pr(θ)Cδ(1r2)\sup_{\theta \in [-\pi,\pi]\setminus (-\delta,\delta)}P_r(\theta)\leq C_{\delta}\cdot (1-r^2) for some constant CδC_\delta depending on δ. Then Pr(θ)P_r(\theta) converges to 0 uniformly when r1r\to 1. Then (π,π)(δ,δ)Pr(θ)dθ0\int_{(-\pi,\pi)\setminus (\delta,\delta)}P_r(\theta)d\theta \to 0 when r1r\to 1.

On the Dirichlet problem

Theorem. Let φ be an integrable function on the unit cirlcle.

(1) uφ(reiθ)=12πππφ(t)1r212rcos(θt)+r2dtu_{\varphi}(re^{i\theta}) = \frac{1}{2\pi} \int_{-\pi}^\pi \varphi(t) \frac{1-r^2}{1-2r\cos(\theta-t)+r^2}dt is a harmonic function on D\mathbb{D}, such that limr1uφ(reiθ)=φ(θ)\lim_{r\to 1}u_\varphi(re^{i\theta}) = \varphi(\theta) when θ is continuous point of φ.

(2) Let φ be a continuous function on the unit circle. Then uφ(reiθ)=12πππφ(t)1r212rcos(θt)+r2dtu_{\varphi}(re^{i\theta}) = \frac{1}{2\pi} \int_{-\pi}^\pi \varphi(t) \frac{1-r^2}{1-2r\cos(\theta-t)+r^2}dt is the unique solution to the Dirichlet problem, and urφu_r\to \varphi uniformly when r1r\to 1.

Proof. Just need to prove uniqueness. Suppose u~\tilde u is another solution, let v=uφu~v = u_\varphi - \tilde u. Then vv is harmonic in D\mathbb{D} and continuous on Dˉ\bar{\mathbb{D}} (the closed unit disk), and v=0v = 0 on the unit circle. Now suppose uφ(z0)u~(z0)u_\varphi(z_0)\neq \tilde{u}(z_0) for some z0Dz_0\in \mathbb{D}, then v(z0)>0|v(z_0)|> 0. By maximum principle, on every C(r)C(r) for z0<r<1|z_0|<r<1, there exists some points wrw_r such that v(wr)>v(z0)|v(w_r)|>|v(z_0)|. Then there must be a subsequence of (wrn)(w_{r_n}) such that wrnwCw_{r_n}\to w'\in C. Since vv is continuous, v(w)v(z0)>0v(w') \geq |v(z_0)|> 0. Contradicts to vv being zero on CC.

Appendix: Heat flow from differential forms

In the note we assume all functions involved are smooth, though most results holds for merely differentiable functions.

Understanding vector fields

A vector field on R3\mathbb{R}^3 is a a vector valued function v(x)=(p(x),q(x),r(x))\mathbf{v}(\mathbf{x}) = (p(\mathbf{x}),q(\mathbf{x}),r(\mathbb{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 CEdn\int_{C}\mathbf{E}\cdot d\mathbf{n}, where n\mathbf n is the outwards normal to the curve CC. Let γ(t)=(x(t),y(t))\gamma(t)=(x(t),y(t)) be a parametrization of the curve CC on [0,1][0,1], then the surface integral is given by 01p(t)y(t)q(t)x(t)dt\int_{0}^1 p(t)y'(t)-q(t)x'(t)dt. We also write it in the form Cpdy+qdx\int_C -pdy+qdx.

On the other hand, if we want to calculate the change of electronic potentials along a curve CC, then it is given by the line integral CEdr=01pdx+qdy\int_{C}\mathbf{E}\cdot d\mathbf{r} = \int_{0}^1 pdx+qdy.

Differential forms on R2\mathbb{R}^2

Here is a breif summery of differential forms in 2-dimensions.

Let DD be a domain with smooth boundary C=DC = \partial D. Then the Green theorem, says Dpdx+qdy=D(qxpy)dxdy\int_{\partial D} p dx + qdy = \int_{D}(\frac{\partial q}{\partial x} - \frac{\partial p}{\partial y})dxdy.

Let ω be the one-form pdx+qdypdx + qdy, then dω=pydydx+qxdxdy=(qxpy)dxdyd\omega = \frac{\partial p}{\partial y} dy\wedge dx + \frac{\partial q}{\partial x}dx \wedge dy = (\frac{\partial q}{\partial x} - \frac{\partial p}{\partial y})dx\wedge dy. Then the Green theorem can be written as Dω=Ddω\int_{\partial D}\omega = \int_{D} d\omega for a one form ω.

Intepreting vector fields

Quite naturally, we can identify the gradient field E=U\mathbf{E} = \nabla U as the 1-form dU=Uxdx+UydydU = \frac{\partial U}{\partial x}dx+\frac{\partial U}{\partial y}dy. Then when we calculate the change of potential along curve CC, it is just given by CdU\int_C dU. Since we have d(dU)=0d(dU) = 0 by direct calculation, when C=DC = \partial D is a closed curve, CdU=Dd(dU)=0\int_C dU = \int_{D}d(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 DD, we do Cpdy+qdx\int_C -pdy+qdx, so it seems we should identify E\mathbf{E} with the 1-form pdy+qdx-pdy+qdx in this case. This motivates us the following operator

Let ω=pdx+qdy\omega = pdx+qdy be a 1-form. Define the *-operator by ω=pdyqdx*\omega = pdy-qdx. Then we have ω=ω**\omega = -\omega.

The * is a spacial case of the Hodge * operator.

In particular,

The Fourier’s law

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\nabla u = -\kappa F, where FF 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 DD is given by Dω\int_{\partial D}\omega, then the Fourier law should be intepreted as ω=κdu\omega = -\kappa *du.

By first law of thermaldynamics, increasing of temperature on D=Dtu(x,t)dm(x)=cDωD = \int_D\frac{\partial }{\partial t}u(\mathbf{x},t)dm\mathbf({x}) = -c \int_{\partial D}\omega. Plug in Fourier law we get Dtu(x,t)dt=κcDddu=κcDΔudxdy\int_{D}\frac{\partial}{\partial t}u(\mathbf{x},t)dt = \kappa c \int_{D}d*du = \kappa c \int_{D}\Delta u dx\wedge dy. Then we get the heat equation.

Polar coordinate

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=uxdx+uydy=urdr+uθdθdu = \frac{\partial u}{\partial x} dx + \frac{\partial u}{\partial y} dy = \frac{\partial u}{\partial r} dr + \frac{\partial u}{\partial \theta}d\theta. And we know that ddu=Δu(x,y)dxdy=Δu(r,θ)rdrdθd*d u = \Delta u(x,y) dx\wedge dy = \Delta u (r,\theta)\cdot rdrd\theta . So it sufficies to calculate * in polar coordinates. From the Cartesian coordinate case, we have dx=dy*dx = dy, dy=dx*dy = -dx, so dx=(dy)=dx**dx = *(dy) = -dx, dy=(dx)=dx=dy**dy = *(-dx) = -*dx=-dy, we conclude that

Lemma. ω=ω**\omega = -\omega.

Note that this is independent of coordinates because we can define the * operator in a coordinate-free way.

Now we calculate ddud*d u in polar coordinates. du=urdr+uθdθdu =\frac{\partial u}{\partial r}dr + \frac{\partial u}{\partial \theta}d\theta, the main question is what is dr*dr and what is dθ*d\theta.

Let’s examine the function f=ln(x2+y2)=lnr2f = \ln(x^2+y^2)=\ln r^2. Then df=frdr+θdθ=2rdr+0df = \frac{\partial f}{\partial r}dr + \frac{\partial }{\partial \theta}d\theta= \frac{2}{r}dr+0. On the other hand, in Cartesian coordinate df=fxdx+fydy=2xx2+y2dx+2yx2+y2dydf = \frac{\partial f}{\partial x}dx + \frac{\partial f}{\partial y}dy = \frac{2x}{x^2+y^2}dx + \frac{2y}{x^2+y^2}dy. df=fydx+fxdy=2rsinθr2d(rcosθ)+2rcosθr2d(rcosθ)=2(cos2θ+sin2θ)dθ=2dθ.*df = -\frac{\partial f}{\partial y}dx + \frac{\partial f}{\partial x}dy = -\frac{2r\sin\theta}{r^2}d(r\cos\theta)+\frac{2r\cos\theta}{r^2}d(r\cos\theta)=2(\cos^2\theta+\sin^2\theta)d\theta = 2d\theta.

We see from this example that it should hold that 2dθ=df=(2rdr)2d\theta = *df = *(\frac{2}{r}dr), so

dr=rdθ.*dr = r d\theta.

Apply =1** = -1 we get dθ=1rdr=1rdr=1rdr*d\theta = *\frac{1}{r}*dr = \frac{1}{r}**dr = -\frac{1}{r}dr.

We are almost done.

Δudxdy=ddu=Δu(r,θ)rdrdθ, \Delta u dx\wedge dy = d*du = \Delta u(r,\theta)rdrd\theta,

In polar coordinates,

ddu=d(urdr+uθdθ)=d(urdr+uθdθ)=d(urrdθ+uθ(1rdr))=(r2ur2+ur)drdθ+(1r2uθ2)dθdr=(2ur2+1rur+1r22θθ2)rdrdθ \begin{split} d*d u &= d*(\frac{\partial u}{\partial r}dr + \frac{\partial u}{\partial \theta}d\theta)\\ &=d(\frac{\partial u}{\partial r}*dr + \frac{\partial u}{\partial \theta}*d\theta)\\ &=d(\frac{\partial u}{\partial r}rd\theta +\frac{\partial u}{\partial \theta}(-\frac{1}{r}dr))\\ &=(r\frac{\partial^2 u}{\partial r^2}+\frac{\partial u}{\partial r})dr\wedge d\theta + (-\frac{1}{r}\frac{\partial^2 u}{\partial \theta^2})d\theta\wedge dr \\ &=(\frac{\partial^2 u}{\partial r^2} +\frac{1}{r} \frac{\partial u}{\partial r}+\frac{1}{r^2}\frac{\partial^2\theta}{\partial\theta^2} )rdr\wedge d\theta \end{split}

We conclude that 2ux2+2uy2=2ur2+1rur+1r22θθ2\frac{\partial^2 u}{\partial x^2}+\frac{\partial^2 u}{\partial y^2} = \frac{\partial^2 u}{\partial r^2} +\frac{1}{r} \frac{\partial u}{\partial r}+\frac{1}{r^2}\frac{\partial^2\theta}{\partial\theta^2}.