Let p>0, a distribution T is p-periodic if τpT=T. Similar to the 1-periodic case, by choosing a partition of unity φn subordinate to the intervals [−np/2,np/2] we get
This time ∑n∈ZFf(x+np)=(ωp(y),Ff(y+x))=(Fτxωp(y),f)=(e−2πixyp1ωp1(y),f(y))=p1∑n∈Zf(pn)e−p2πinx. Of course, one can also derive this by applying the Poisson summation formula directly.
This implies FT=∑n∈Z(T(y),φ0e−p2πiny)p1δpn. If we write T^(n)=p1(T(y),φ0e−p2πiny), then by taking inverse transform (recall that F−1δpn=ep2πinx), we can write T=∑n∈ZT^(n)ep2πinx, this is the Fourier expansion of T, which is parallel with the Fourier series theory.
Let D:=−idxd be the Dirac operator and s be a complex number.
In problem set 3 we gave solution of the equation (D+s)f=g by explicitedly calculating the inverse of D+s using Fourier series and showed that the solution is given by f=g∗G, where G is given by Fourier series ∑n∈Zn+s1einx.
But this method may not work for general differential operators. In this note we’ll revisit the problem using common method of solving differential equations from distribution theory. From the derivative theorem we see that (D+s)G has Fourier coefficients 1, so as a 2π-periodic distribution, (D+s)G is equal to 2πω2π. From the above point of view, we can interprate G as a fundamental solution of 2π-periodic functions since ω2π is the unit of convolution for 2π-periodic functions. (Recall that convolving with ω2π = periodization to period 2π, but if it’s already 2π-periodic then periodization does nothing). You may wonder why (D+s)=2πω2π instead of ω2π, this is because when we defined convolution at that time, we made a normalization by deviding 2π.
One of the applications of Fourier transform is that it can help us to find solution of partial differential equations in a direct way. As an example we consider the simple case D+s. Later we’ll use the Fourier transform to solve the wave equation and heat equation.
Let D be an open set with C1 bundary on C. We adapt the identification C with R2 via z with x+iy, where i=−1.
Then for f(z) which is C1 on a neighbourhood of D, we have the following Cauchy integral formula
In the real notation it is f(ζ)=2πi1∫∂Dz−ζf(z)dz−π1∫Dz−ζ∂zˉf(z)dλ(z), where dλ(z)=dxdy stands for the Lebesgue measure on C=R2.
Proof. We give a proof of this using a distribution version of divergence theorem (or Green formula in 2 dimensional case), see Problem set 8 problem 3. The result is
We’ll use the complex version of this result. Choose a parametrization γ(t)=(x(t),y(t)) of ∂D, such that the tangent is given by (x′(t),y′(t)) (or γ′(t)=x′(t)+iy′(t)) and invards normal is given by (−y′(t),x′(t)) (or iγ′(t)=i(x′(t)+iy′(t))=−y′(t)+ix′(t)). Then ds=∣γ′(t)∣dt, n=∣γ′(t)∣1(−y′(t),x′(t)) so n1ds=−y′(t)dt and n2ds=x′(t)dt. Then (−∂x[χD],f)=∫D∂x∂fdxdy=−(n1ds,f)=∫f(x(t),y(t))y′(t)dt=∫∂Dfdy. Similarly (−∂y[χD],f)=−∫∂Dfdx. Then we get
or ∂zˉ[χD]=2i1dz∣∂D.
One can easily translate the proof using green formula or Stokes theorem. Let χD be the characteristic function of D, χBϵ be the characteristic function of a small ball of radius ε centered at ζ.
Apply the identity (2) to D∖Bϵ and acts to the function z−ζf(z) we get ∫∂Dz−ζf(z)dz−∫∂Bϵz−ζf(z)dz=2i(∫Dz−ζ∂zˉfdxdy−∫Bϵz−ζ∂zˉfdxdy). Note that by parametrizing ∂Bϵ with γ(t)=ϵeit+ζ, we get ∫∂Bϵz−ζf(z)dz=i∫02πf(ζ+ϵeiθ)dθ which goes to 2πf(ζ) when ϵ→0, by dominated convergence theorem.
Since ∫Bϵ∣z−ζ∂zˉf∣dxdy≤supBϵ∣∂zf∣⋅∫Bϵr1rdrdθ=2πϵsupBϵ∣∂zˉf∣→0 when ϵ→0, we get 2πif(ζ)=∫∂Dz−ζf(z)dz−2i∫Dz−ζ∂zˉfdxdy. From the notation dz∧dzˉ=(dx+idy)∧(dx−idy)=−2idx∧dy, we get f(ζ)=2πi1(∫∂Dz−ζf(z)dz+∫Dz−ζ∂zˉf(z)dz∧dzˉ).
Proof.
Let φ be a smooth function of compact support on C. Let D be a domain with C1 boundary which contains the support of φ. Then by Cauchy integral formula φ(ζ)=2πi1∫∂Dz−ζφ(z)dz−π1∫Dz−ζ∂zˉφ(z)dλ(z)=0−π1∫Dz−ζ∂zˉφdλ(z) because the support of φ is inside D and hence φ≡0 on ∂D.
It follows that
(∂zˉ[π1z−ζ1],φ)=−π1∫Dz−ζ∂zˉφdλ(z)=φ(ζ)=(δζ,φ).