Lecture 17

The Fourier-Laplace transform

Let gg be a timelimted smooth function, say suppg[R,R]\text{supp} g\subset [-R,R]. Then the Fourier transform of gg can be extended to a complex valued function Fg(z):=+g(t)e2πiztdt\mathcal Fg(z):=\int_{-\infty}^{+\infty}g(t)e^{-2\pi i z t}dt. Let z=x+iyz = x+iy, then Fg(z)=RRg(t)e2πitxe2πytdt\mathcal Fg(z) = \int_{-R}^{R}g(t)e^{-2\pi i tx}e^{2\pi yt}dt. Moreover, zˉFg(z)=RRg(t)zˉe2πitzdt=0\partial_{\bar z}\mathcal Fg(z) = \int_{-R}^R g(t)\partial_{\bar z}e^{-2\pi i t z}dt = 0 because e2πitze^{-2\pi i t z} is holomorphic. This implies Fg(z)\mathcal Fg(z) is a holomorphic function on the whole complex plane, we call such function an entire function.

Another way to see that Fg\mathcal F g is entire is to use power series. From complex analysis we know that ez=k0znk!e^{z} = \sum_{k\geq 0}\frac{z^n}{k!} and the series converges unformly on compact sets of C\mathbb{C}, so e2πizt=k0(2πizt)kk!=k0(2πi)ktk1k!zke^{-2\pi i zt} = \sum_{k\geq 0 }\frac{(-2\pi i z t)^k}{k!} = \sum_{k\geq 0}(-2\pi i)^kt^k\frac{1}{k!}z^k. If gg is supported on [R,R][-R,R], then RRg(t)e2πitzdt=RRk0(2πiz)kk!tkg(t)dtsupgk02Rk+1(2πi)kzkk!\int_{-R}^{R}g(t)e^{-2\pi i t z}dt =\int_{-R}^R \sum_{k\geq 0}\frac{(-2\pi i z)^k}{k!} t^k g(t)dt \leq \sup|g|\cdot\sum_{k\geq 0} 2R^{k+1}(-2\pi i)^k\frac{z^k}{k!}. The Cauchy-Hadamard criterion implies that the series on the right hand side has convergence radius and defines an entire function. To justify the switch of integral and summation, let sn=k=0n(2πiz)kk!tks_n = \sum_{k=0}^n\frac{(-2\pi i z)^k}{k!} t^k , then snk0(2πtz)kk!=e2πzt|s_n|\leq \sum_{k\geq 0}\frac{(2\pi|t||z|)^k}{k!} = e^{2\pi |z||t|} which is continuous and thus integrable on [R,R][-R,R] (as a function of tt) and dominate sn|s_n|, then apply dominated convergence theorem.

We’ll consider the power series expansion in the proof of central limit theorem in later lecture.

Paley-Wiener theorem

Theorem

Let U(z)U(z) be an entire function with growth condition

(1+z)kU(z)Cke2πRIm(z) (1+|z|)^k|U(z)|\leq C_ke^{2\pi R|Im(z)|}

(eqn:plcondition) for every k0k\geq 0. Then UU is the Fourier-Laplace theorem of a smooth function supported on [R,R][-R,R].

Proof. The idea to find uu is Fourier inversion. Let U(x)U(x) be the restriction of UU on the real line. Then the growth condition eqn:plcondition implies that U(x)U(x) is a Schwartz function. To see this, first the inequality implies U(x)U(x) is rapidly decreasing, then by Cauchy estimate, the derivative of a holomorphic function can be estimated by its value. Then u(t)=RxU(x)e2πixtdxu(t) = \int_{\mathbb{R}_x}U(x)e^{2\pi i x t}dx is a Schwartz function. To show that UU is the Fourier Laplace transform of uu, we’ll show that uu is supported in [R,R][-R,R]. Then Fu(z)=U(z)\mathcal Fu(z) = U(z) because they are entire functions but coincide on the real line.

We claim that u(t)=+U(x)e2πixtdx=0u(t) = \int_{-\infty}^{+\infty}U(x)e^{2\pi i x t}dx = 0 when t>R|t|>R. For r>0r>0 and bRb\in \mathbb{R} (note: we are not assuming b>0b>0 even though the picture looks like b>0b>0), let I1,I2,I3,I4I_1,I_2,I_3,I_4 be path integrals as shown in the picture, they depend on rr though not explicitly written. Since U(z)U(z) is entire function, the Cauchy integral formula (or Morera theorem) implies I1+I2+I3+I4=0I_1+I_2+I_3+I_4 = 0. Note that limrI1=u(t)\lim_{r\to \infty}I_1 = u(t). First, we show that I2I_2 and I4I_4 go to zero when rr\to \infty, so we get u(t)=limrI3u(t) = -\lim_{r\to \infty}I_3.

When b>0b>0, I20bU(r+iy)e2πit(r+iy)dysupy[0,b]U(r+ib)0be2πtydy|I_2|\leq \int_{0}^b |U(r+iy)||e^{2\pi i t(r+iy)}|dy \leq \sup_{y\in [0,b]}|U(r+ib)|\int_{0}^b e^{-2\pi t y}dy. The growth condition on U(z)|U(z)| implies supy[0,b]U(r+ib)Cke2πRb(1+r2+b2)k\sup_{y\in [0,b]}|U(r+ib)|\leq C_k \frac{e^{2\pi R|b|}}{(1+\sqrt{r^2+b^2})^k}. So I2conste2πRb(1+r2+b2)k0|I_2|\leq const\cdot \frac{e^{2\pi R|b|}}{(1+\sqrt{r^2+b^2})^k}\to 0 when rr\to \infty. The same result holds when b<0b<0 and for I4I_4 by the same argument.

We have showed that for every bRb\in \mathbb{R}, u(t)=limrI3=U(x+ib)e2πt(x+ib)dxu(t) = -\lim_{r\to \infty}I_3 = \int_{-\infty}^{\infty}U(x+ib)e^{2\pi t(x+ib)}dx, so u(t)C2e2πRb2πtb+1(1+x)2dx|u(t)|\leq C_2 e^{2\pi Rb - 2\pi t b} \int_{-\infty}^{+\infty} \frac{1}{(1+|x|)^2}dx for every bb. When b>0b>0, 2πRb2πtb<02\pi Rb - 2\pi t b<0 when Rt<0R-t<0, i.e. t>Rt>R, this implies u(t)=0u(t)=0 because the non-negative number which is less than any negative exponential is 0. Similarly, when b<0b<0, we 2πRb2πtb<02\pi Rb - 2\pi tb<0 when t<Rt<-R, which implies u(t)=0u(t) = 0 when t<Rt<-R. The proof is complete.

Distribution with compact support

The Paley-Wiener theorem can be generalized to distirbutions. But

Let TT be a distribution. We say TT is zero or zero mass at xx if for some open neighbourhood UU of xx, (T,φ)=0(T,\varphi) = 0 for every φCc(U)\varphi\in \mathscr{C}_c^{\infty}(U). As the case of functions, the support of TT is the set where TT is not zero, which is automatically closed. The set E\mathscr{E}' denote the distributions with compact support.

One important classes of distribution with compact support are compact supported measures.

Topology of E\mathscr{E}'

Recall that a distribution is a coninuous linear functional on Cc(R)\mathscr{C}_c^{\infty}(\mathbb R). A natural question is the dual of C(R)\mathscr{C}^{\infty}(\mathbb{R}). For a distribution TT, (T,f)(T,f) may not be defined when fC(R)f\in \mathscr{C}^\infty(\mathbb{R}) even when TT is a continuous function. Not suprisingly this can be done when TT has compact support.

Let TT be a distribution supported on KK, where KK is a compact set. Let ψCc(R)\psi \in \mathscr{C}_c^{\infty}(R) which is 1\equiv 1 on KK. For fC(R)f\in \mathscr{C}^{\infty}(\mathbb{R}), we write f=ψf+(1ψ)ff = \psi f + (1-\psi)f. The (T,ψf)(T,\psi f) is defined because ψf\psi f has compact support, the other part (1ψf)(1-\psi f), though may not have compact support, is gurenteed to support outside the support of TT so it’s meaningful to assign zero. So we define (T,f):=(T,ψf)(T,f):=(T,\psi f). To show that it is well-defined, let ψCc(R)\psi'\in \mathscr{C}_c^{\infty}(\mathbb{R}) satisfying also ψ1\psi' \equiv 1 on KK. Then (T,ψf)(T,ψf)=(T,(ψψ)f)=0(T,\psi f) - (T,\psi' f) = (T,(\psi - \psi')f) = 0 because ψψ\psi - \psi' is supported outside KK. Moreover we see that (T,f)=(T,ψf)CK0lksupxK(fψ)(l)(x)CKlksupxKf(l)(x)|(T,f)| = |(T,\psi f)|\leq C_K \sum_{0\leq l\leq k}\sup_{x\in K}|(f\psi)^{(l)}(x)|\leq C_K' \sum_{l\leq k}\sup_{x\in K}|f^{(l)}(x)| for every fC(R)f\in \mathscr{C}^{\infty}(\mathbb{R}).

Note that the topology of C(R)\mathscr{C}^{\infty}(\mathbb{R}) is determined by the semi-norms l,K\|\cdot\|_{l,K}, and fnff_n\to f if and only if for every compact KK and every ll, fnfl,K0\|f_n - f\|_{l,K}\to 0. Then for TET\in \mathscr{E}', (T,fn)(T,f)CKTlkfnfl,KT0|(T,f_n)-(T,f)|\leq C_{K_{T}}\cdot \sum_{l\leq k}\|f_n - f\|_{l,K_T}\to 0. If we denote by E\mathscr{E} the space C(R)\mathscr{C}^{\infty}(\mathbb{R}) together with its topology, then E\mathscr{E}' are continuous linear functionals on E\mathscr{E}, which justifies the notation.

Action of distribution depending on parameters

To verify the Cauchy-Riemann Fourier-Laplace transform of distribution we need to “differenciate under action of distributions”. These nice properties are consequence of continunity of TT, as we’ll see.

(=lem:diff)

Lemma

Let ϕ(x,y)C(X×Y)\phi(x,y)\in \mathscr{C}^{\infty}(X\times Y), where X,YX,Y are open sets on R\mathbb{R}. Let TET\in \mathscr{E}'. Then

(1) (T(y),ϕ(x,y))(T(y),\phi(x,y)) defines a smooth function on XX.

(2) x(k)(T(y),ϕ(x,y))=(T(y),xk(ϕ(x,y)))\partial^{(k)}_x(T(y),\phi(x,y)) = (T(y),\partial^{k}_x(\phi(x,y))).

Proof.

Let g(x)=(T(y),ϕ(x,y))g(x) = (T(y),\phi(x,y)). First we show that g(x)g(x) is continuous. Note that if ϕ(x,y)C(X,Y)\phi(x,y)\in \mathscr{C}^{\infty}(X,Y), then ϕ(x+h,y)ϕ(x,y)\phi(x+h,y)\to \phi(x,y) in C(Y)\mathscr{C}^{\infty}(Y) when h0h\to 0. Since TT is compact support, TT is continuous on C(Y)\mathscr{C}^\infty(Y), this implies g(x+h)=(T(y),ϕ(x+h,y))(T(y),ϕ(x,y))=g(x)g(x+h) = (T(y),\phi(x+h,y))\to (T(y),\phi(x,y)) = g(x) when h0h\to 0.

Then we consider derivatives using this observation. Taylor expansion of ϕ(x,y)\phi(x,y) near xx implies ϕ(x+h,y)=ϕ(x,y)+hxϕ(x,y)+h22xxϕ(x+θh,y)\phi(x+h,y) = \phi(x,y)+h\partial_x\phi(x,y) + \frac{h^2}{2}\partial_{xx}\phi(x+\theta h,y) for θ<1|\theta|<1. Then g(x+h)=g(x)+h(T,xϕ)+h22(T,xxϕ(x+θh,y))g(x+h) = g(x)+h (T,\partial_x\phi)+\frac{h^2}{2}(T,\partial_{xx}\phi(x+\theta h,y)).

g(x+h)g(x)h(T,xϕ)=h2(T(y),xxϕ(x+θh),y)h2(T(y),xxϕ(x,y)+ϵ)0|\frac{g(x+h)-g(x)}{h}-(T,\partial_x\phi)| = \frac{h}{2}|(T(y),\partial_{xx}\phi(x+\theta h),y)|\leq \frac{h}{2}(|T(y),\partial_{xx}\phi(x,y)|+\epsilon)\to 0 when h0h\to 0. The ε comes continunity of (T,xx(x,y)(T,\partial_{xx}(x,y) when hh is small.

Iterating the above argument shows that gg is smooth and its derivative is given by g(k)(x)=(T(y),x(k)ϕ(x,y))g^{(k)}(x) = (T(y),\partial_x^{(k)}\phi(x,y)).

Obtaining smooth function from distribution

In problem set 7, we proved the following result

The second bullet is a consequence of C\mathscr{C}^\infty convergence of Riemann sum, still by a continunity argument. To verify this, we must show for every gCc(R)g\in \mathscr{C}_c^{\infty}(\mathbb{R}), (ϕT,g)=(T,ϕg)=(Tϕ)(x)g(x)dx(\phi*T,g) = (T,\phi^-*g) = \int (T*\phi)(x)g(x)dx. Let Rh=kZϕ(xkh)g(kh)hR_h = \sum_{k\in \mathbb{Z}}\phi^-(x-kh) g(kh)h be the Riemann sum of ϕg\phi^-*g. (T,Rh)=kZhg(kh)(T,ϕ(xkh))=kZhg(kh)(Tϕ)(kh)(T,R_h) = \sum_{k\in \mathbb{Z}}h g(kh)(T,\phi^-(x - kh)) = \sum_{k\in \mathbb{Z}}hg(kh)(T*\phi)(kh). Since RhϕgR_h\to \phi^-*g in Cc(R)\mathscr{C}_c^\infty(\mathbb{R}) and TT is a distribution, the LHS \to (T,ϕg)=(ϕT,g)(T,\phi^-*g) = (\phi*T,g) when h0h\to 0. But observe that the RHS, as a Riemann sum of (Tϕ)(x)g(x)dx\int (T*\phi)(x)g(x)dx, goes to (Tϕ)(x)g(x)dx\int (T*\phi)(x)g(x)dx. This finishes the proof. Note that actually the summation kZ\sum_{k\in \mathbb{Z}} is actually a finite sum because ϕ,ψ,g\phi,\psi,g are compactly supported.

Fourier-Laplace transform of distributions with compact support

Alternative definition of Fourier transform for compactly supported distributions

Recall that when we defined the Fourier transform for distributions, we tried a function (FTf,g)=yxf(x)e2πixydxdy(\mathcal F T_f,g) = \int_y \int_x f(x)e^{-2\pi i xy}dx dy and applied Fubini theorem to conclude that (FTf,g)=(T,Fg)(\mathcal FT_f,g) = (T,\mathcal Fg) and defined the Fourier transform to be (FT,g)=(T,Fg)(\mathcal FT,g) = (T,\mathcal Fg). But what if we do not use Fubini theorem, and then (FTf,g)=((Tf(x),e2πixy),g(y))(\mathcal FT_f,g) = ((T_f(x),e^{-2\pi i x y}),g(y)), then try to define FT\mathcal FT be the function y(T(x),e2πixy)y\mapsto (T(x),e^{-2\pi i xy})? Unfortunately this would not be successful for tempered distribution TT because e2πixye^{-2\pi i xy} is not compacted supported and this function can be infinity. But if TT has compact support, then TT acts on all smooth functions and this approach works! Moreover, by the continunity argument, the function (T(x),e2πixy)(T(x),e^{-2\pi i x y}) is a smooth function on yy.

Definition. Let TT be a distribution with compact support. The Fourier transform of TT is a smooth function given by FT(x)=(T(y),e2πixy)\mathcal FT(x) = (T(y),e^{-2\pi i xy}).

Of course, we need to show the new definition of Fourier transform as a smooth function concide (in the sense of distribution) as the original definition. By the identity principle of distributions, it sufficies to check that their actions on Cc\mathscr{C}_c^{\infty} are the same, i.e. y(T(y),e2πixy)g(x)dx=(T,Fg)\int_{y}(T(y),e^{-2\pi i xy})g(x)dx = (T,\mathcal Fg). This follows from the same argument of proving Tϕ=ϕTT*\phi = \phi*T by writing Fg=g(y)e2πixy\mathcal Fg = \int g(y)e^{-2\pi i x y} as a Riemann sum kZg(kh)e2πixkhh\sum_{k\in \mathbb{Z}}g(kh)e^{-2\pi i xkh}h.

The Fourier-Laplace transform

Now we’re ready to define the Fourier-Laplace transform of compactly supported distributions.

By the “differentiation under integral sign” lemma, zˉFT(z)=(T(t),zˉe2πitz)=0\partial_{\bar{z}}\mathcal FT(z) = (T(t),\partial_{\bar{z}}e^{-2\pi i t z}) = 0, and FT(z)=0\mathcal FT(z) = 0.

Example

The Fourier-Laplace transform of δa\delta_a is Fδa(z)=(δa(t),e2πizt)=e2πiaz\mathcal F\delta_a(z) = (\delta_a(t),e^{-2\pi i zt}) = e^{-2\pi i a z}. In particular, F(δa+δa2)=cos(2πaz)\mathcal F(\frac{\delta_a + \delta_{-a}}{2}) = \cos(2\pi a z), F(δa+δa)=2isin(2πaz)\mathcal F(-\delta_a + \delta_{-a}) = 2i\sin(2\pi a z).

Let χ[a,a]\chi_{[-a,a]} be the step function. Then Fχ[a,a](z)=(χa,a(t),e2πitz)=aae2πitzdt=12πiz(e2πiaze2πiaz)=2isin(2πaz)2πiz=sin(2πaz)πz\mathcal F\chi_{[-a,a]}(z) = (\chi_{-a,a}(t),e^{-2\pi i t z}) = \int_{-a}^a e^{-2\pi i t z}dt = \frac{1}{-2\pi i z}(e^{-2\pi i az}-e^{2\pi i az}) =\frac{-2i \sin(2\pi a z)}{-2\pi i z} = \frac{\sin(2\pi a z)}{\pi z}. This is an entire function with power series 1z(zz33!+z55!+)=n=0+(1)nz2n(2n+1)!\frac{1}{z}\cdot (z -\frac{z^3}{3!}+\frac{z^5}{5!}+\cdots) = \sum_{n = 0}^{+\infty}\frac{(-1)^nz^{2n}}{(2n+1)!}.