Lecture 2

4. Heat equation on the circle

Derivation of heat equation

In this note we identify the circle S1S^1 with T:=R/Z\mathbb{T}:=\mathbb{R}/\mathbb{Z} via e2πixe^{2\pi i x} \leftrightarrow x[0,1]x\in [0,1]. When we say the point xx on the circle, we mean the point e2πixS1e^{2\pi i x}\in S^1. In our discussion “time” is a real variable t[0,+]t\in [0,+\infty].

The distribution of temperature on the circle is modeled by a density function f(x)f(x) for xTx\in \mathbb{T}. The temperature on interval [a,b][a,b] is given by abf(x)dx\int_{a}^b f(x)dx.

Let u(x,t)u(x,t) be the temperature distribution of circle at time tt, note that u(0,t)=u(1,t)u(0,t)=u(1,t), so the temperature of an interval (a,b)(a,b) is given by abu(x,t)dx\int_{a}^b u(x,t)dx. Let u(x,0)=f(x)u(x,0)=f(x) be the initial temperature. We study the heat diffusion when time increases under the following postulates:

  1. Increasing of temperature of an interval (a,b)(a,b) during time Δt\Delta t is proportional to increasing of heat on (a,b)(a,b). In other words, [T(a,b,t+ϵ)T(a,b,t)]=c[H(a,b,t+ϵ)H(a,b,t)][ T(a,b,t+\epsilon)-T(a,b,t)] = c\cdot [H(a,b,t+\epsilon)-H(a,b,t)] for some constant c>0c>0, or tH(a,b,t)=ctT(a,b,t)\frac{\partial}{\partial t}H(a,b,t) = c\frac{\partial}{\partial t}T(a,b,t) if we look at infinitesimal change of time tt.

  2. Fix time tt, consider heat transfer during an infinitesimal time interval. Let ωt(x)\omega_t(x) be the flux of heat at xx, i.e. the amount of heat which goes from leftside to right side at xx, so that ωt(x)\omega_t(x) is positive if heat transfers from left to right and negative if heat transfers from right to left. We get

    tH(a,b,t)=ωt(a)ωt(b).\frac{\partial}{\partial t}H(a,b,t) = \omega_t(a)-\omega_t(b).
  3. Fourier’s law: the heat flux is negatively proportional to the derivative of temperature, i.e. at time tt, ω(x)=κlimδ0u(x+δ,t)u(x,t)δ=κxu(x,t)\omega(x) = \kappa \lim_{\delta\to 0}\frac{u(x+\delta,t)-u(x,t)}{\delta} = -\kappa \frac{\partial}{\partial x}u(x,t) for a constant κ>0\kappa > 0. Note that negative because heat transfers from higher temperature to lower temperature.

    • For example, if the temperature is higher on the right, so u(x,t) is increasing in x variable, so the partial derivative \frac{\partial}{\partial x}u(x,t) is positive, but in this case the heat will go from right(higher) to the left(lower), so \omega_t(x) < 0 because \omega_t(x) means the amount of heat going from left to right.

    Consequently, xωt(x)=κ2x2u(x,t)\frac{\partial}{\partial x}\omega_t(x) = -\kappa \frac{\partial^2}{\partial x^2}u(x,t).

From the first two postulates, we get ddtabu(x,t)dx=abtu(x,t)dx=c(ωt(a)ωt(b))=cabωt(x)dx\frac{d}{dt}\int_{a}^bu(x,t)dx = \int_{a}^b \frac{\partial}{\partial t}u(x,t)dx = c(\omega_t(a)-\omega_t(b)) = -c\int_{a}^b\omega_t'(x)dx. By the third postulate, ωt(x)=κ2x2u(x,t)\omega_t'(x)=-\kappa \frac{\partial^2}{\partial x^2}u(x,t), so we get abtu(x,t)dx=ab2x2cκu(x,t)dx\int_{a}^b\frac{\partial}{\partial t}u(x,t)dx = \int_{a}^b \frac{\partial^2}{\partial x^2}c\kappa u(x,t)dx. Equality holds for any interval so we get the equation

tu(x,t)=cκx2u(x,t).\frac{\partial}{\partial t}u(x,t) = c\kappa \frac{\partial}{\partial x^2}u(x,t).

We consider the standard form ut=12uxxu_t = \frac{1}{2}u_{xx}.

Solving the heat equation using Fourier series

Since u(x,t)u(x,t) is periodic on the xx variable, Fourier’s idea is to look for solutions of the form u(x,t)=n=+cn(t)e2πinxu(x,t) = \sum_{n=-\infty}^{+\infty}c_n(t)e^{2\pi i n x}. It follows that cn(t)=01u(x,t)e2πinxdxc_n(t) = \int_{0}^1 u(x,t)e^{-2\pi i n x}dx. Take ddt\frac{d}{dt} on cn(t)c_n(t) we get

cn(t)=01tu(x,t)e2πinxdx=1201uxx(x,t)e2πinxdx.c_n'(t) = \int_{0}^1 \frac{\partial}{\partial t}u(x,t)e^{-2\pi i n x}dx = \frac{1}{2}\int_{0}^1 u_{xx}(x,t)e^{-2\pi i n x}dx.

(It makes sense to assume uu to be smoothly depending on tt so that we can change derivative and integral here, by applying mean value theorem).

By the derivative theorem, g^(n)=2πing^(n)\widehat{g'}(n) = 2\pi i n \hat{g}(n), so apply to u(x,t)u(x,t) in the xx variable we get cn(t)=2π2n2cn(t)c_n'(t) = -2\pi^2n^2c_n(t). Solving the ODE we get cn(t)=cn(0)e2π2n2tc_n(t) = c_n(0)e^{-2\pi^2n^2 t}. cn(0)c_n(0) is the Fourier coefficient of u(x,0)=f(x)u(x,0) = f(x), so we get u(x,t)=n=+f^(n)e2π2n2te2πinxu(x,t) = \sum_{n = -\infty}^{+\infty}\hat{f}(n)e^{-2\pi^2n^2 t}e^{2\pi i n x}.

Write back f^(n)=01f(y)e2πinydy\hat{f}(n) = \int_{0}^1 f(y)e^{-2\pi i n y}dy, then u(x,t)=n=+01f(y)e2πinye2πinxe2π2n2tdy=01f(y)e2π2n2te2πin(xy)dyu(x,t) = \sum_{n = -\infty}^{+\infty} \int_{0}^1 f(y)e^{-2\pi i n y}e^{2\pi i n x}e^{-2\pi^2n^2 t}dy = \int_{0}^1 f(y)e^{-2\pi^2n^2 t}e^{2\pi i n (x-y)}dy, provided we can change the integral and summation. Let g(x,t)=n=+e2π2n2te2πinxg(x,t) = \sum_{n=-\infty}^{+\infty}e^{-2\pi^2n^2 t}e^{2\pi in x}, then the series converges absolutely hence uniformly in xx variable for every tt, (e2πn2t<\sum e^{-2\pi n^2 t}<\infty as a geometric series), then g(x,t)g(x,t) is a continuous function on the circle, this also justifies the change of integral and summation.

We obtained a simpler form of solution: u(x,t)=01g(xy,t)f(y)dyu(x,t) = \int_{0}^1g(x-y,t)f(y)dy. This type of integrals is called convolution, as we shall study next.

5. Convolution

Definition 5.1. Let f,gf,g be integrable periodic functions with period 1. The convolution of f,gf,g is a function determined pointwise by

(fg)(x)=01f(xy)g(y)dy (f*g)(x) = \int_{0}^1f(x-y)g(y)dy

Remarks.

  • By changing variable y=xyy' = x-y, we get 01f(xy)g(y)dy=xx1f(y)g(xy)d(y)=01f(y)g(xy)dy=01g(xy)f(y)dy\int_{0}^1 f(x-y)g(y)dy = \int_{x}^{x-1}f(y')g(x-y')d(-y') = -\int_{0}^{-1}f(y')g(x-y')dy' = \int_{0}^1g(x-y)f(y)dy, so we can also define convolution by (fg)(x)=01g(xy)f(y)dy(f*g)(x) = \int_{0}^1 g(x-y)f(y)dy.

  • Our definition of convolution here is periodic convolution. Since convolution means taking average, we need to normalize in period TT so that fg=1T0Tf(xy)g(y)dyf*g = \frac{1}{T}\int_{0}^Tf(x-y)g(y)dy. The standard notion of convolution is on the real line and plays an important rule in Fourier transforms.

  • The NN-th Fourier partial sum SN(f)S_N(f) can be written as fDNf*D_N. In fact, SN(f)(x)=n=NNf^(n)e2πinx=n=NN01f(y)e2πinydye2πinx=01f(y)n=NNe2πin(xy)dy=fDN(x)S_N(f)(x) = \sum_{n=-N}^N \hat{f}(n)e^{2\pi inx} = \sum_{n=-N}^N \int_{0}^1f(y)e^{-2\pi i n y} dy \cdot e^{2\pi i n x} =\int_{0}^1f(y) \sum_{n = -N}^N e^{2\pi i n (x-y)} dy = f*D_N(x), here DN=n=NNe2πinxD_N = \sum_{n=-N}^N e^{2\pi i n x} is the Dirichlet kernel on T\mathbb{T}.

Properties of convolution

Let f,g,hf,g, h be integrable periodic functions with period 1.

Linearity

The linearity of convolutions follows directly from linearity of integrals.

  • (f+g)h=fh+gh(f+g)*h = f*h+g*h.
  • (cf)g=cfg(cf)*g = cf*g.

Commutativity & Associativity

  • fg=gff*g = g*f. This is just the first remark of Definition 5.1.
  • (fg)h=f(gh)(f*g)*h = f*(g*h). This follows from the Fubini theorem.

From the first properties one may feel that convolution looks quite similar to multiplication. This intuition may be justified in some sense by the fact that convolution and multiplication are Fourier transform of each other.

Convolution theorem

Let f,gf,g be integrable period 1, then fg^(n)=f^(n)g^(n)\widehat{f*g}(n) = \hat{f}(n)\hat{g}(n).

Proof. By definition of convolution and Fubini theorem, fg^(n)=01(fg)(x)e2πinxdx=xyf(xy)g(y)dye2πinxdx=yxf(xy)e2πin(xy)dx  e2πinyg(y)dy=f^(n)g^(n)\widehat{f*g}(n) = \int_{0}^1 (f*g)(x)e^{-2\pi i n x}dx = \int_x\int_y f(x-y)g(y)dy e^{-2\pi i nx}dx = \int_{y}\int_xf(x-y)e^{-2\pi i n(x-y)}dx \; e^{-2\pi i n y }g(y)dy = \hat{f}(n)\hat{g}(n).

  • The integrand satisfies |f(x-y)g(y)e^{-2\pi i n x}| = |f(x-y)g(y)| and \int_y\int_x |f(x-y)||g(y)|dxdy = \int_y |g(y)|\int_x |f(x)|dx dy = \|f\|_{L^1}\|g\|_{L^1} < \infty because f,g are both L^1 on [0,1] .

Continuty of convolution

Another good property of convolution is its regulartiy. The convolution is defined by integration, which are good because integrals usually have more regularity than integrands. In other words, fgf*g is “nicer” than the nicer one of ff and gg, that’s why convolution is widely used to do approximations. As an example we present the approximation of integrable functions in detail to let you have some feeling about this technique.

Proposition 5.1. Let f,gf,g be periodic functions with period 1. Then

  • If ff is continuous, gg is integrable, then fgf*g is continuous.
  • Moreover, if ff is integrable, gg is bounded, then fgf*g is also continuous.

Proof. Let hh be a real number. fg(x+h)fg(x)=01(f(x+hy)f(xy))g(y)dy01f(x+hy)f(xy)g(y)dy|f*g(x+h) - f*g(x)| = |\int_{0}^1 (f(x+h-y)-f(x-y))g(y)dy|\leq \int_{0}^1|f(x+h-y)-f(x-y)||g(y)|dy. For ϵ>0\epsilon > 0, since ff is continuous, when h|h| is small enough, f(x+h)f(x)<ϵ|f(x'+h)-f(x')|<\epsilon for every xRx'\in \mathbb{R}. It follows that f(x+hy)f(xy)<ϵ|f(x+h-y)-f(x-y)|<\epsilon for every y(0,1)y\in (0,1). Then the right hand side ϵgL1\leq \epsilon \|g\|_{L^1}, where gL1:=01g(y)dy<\|g\|_{L^1}:=\int_{0}^1|g(y)|dy<\infty because gg is integrable. This shows that fgf*g is continuous. Actually fgf*g is uniformly continuous in this case.

To prove the second stronger assertion we need an approximation lemma which will be proved in problem set 1.

Lemma 5.2. Let ff be an integrable function on [0,1][0,1], then for ϵ>0\epsilon > 0, there exists a continuous function f~\tilde f such that ff~L1<ϵ\|f - \tilde f\|_{L^1}< \epsilon.

By the approximation lemma, choose a sequence of continuous functions fkf_k such that ffkL10\|f- f_k\|_{L^1} \to 0 when kk\to \infty. Then fg(x)fkg(x)=(ffk)g(x)=01(ffk)(y)g(xy)dyBffkL1|f*g(x) - f_k*g(x)| = |(f-f_k)*g(x)|=\int_{0}^1|(f - f_k)(y)||g(x-y)|dy\leq B\|f-f_k\|_{L^1}, where 0<B<0<B<\infty is an upper bound of g|g|. Since the right hand side goes to 0 when kk\to \infty and does not depend on xx, we conclude that fkgfgf_k*g\to f*g uniformly. Since fkgf_k*g is continuous because fkf_k is continuous, this implies that fgf*g is a uniform limit of continuous functions, which is continuous.

The approximation theorem

Here are the idea to prove Lemma 5.2 for the Riemann intregral case. First we give some simple notions to shorten the discussion.

Definition 5.3. Let AA be a set. Let χA\chi_A given by

χA(x)={1,xA0,xA \chi_A(x) = \begin{cases}1,x\in A \\ 0,x\notin A\end{cases}

Definition 5.4. A rect function is a function of the form χ(a,b)\chi_{(a,b)}, where (a,b)(a,b) is an interval. rect stands for rectangle, the graph of a rect function looks like a rectangle hence the name.

Definition 5.5. A step function is given by k=1mckχ(ak,bk)\sum_{k=1}^m c_k\chi_{(a_k,b_k)}, where ckc_k are real numbers, (ak,bk)(a_k,b_k) are disjoint intervals. Still, the name explains its shape. A step function is a finite linear combination of rect functions of disjoint intervals.

Now let ff be a Riemann integrable function.

Step 0. By definition, there exists a step function hh such that fhL1<ϵ\|f - h\|_{L^1}<\epsilon. In fact, this is how the Riemann integral is defined. The integral of a function is defined by taking limit of integral of step functions for suitable functions for which the limit makes sense.

Step 1. Approximate step function by continuous functions. That is, find a continuous function gg such that ghL1<ϵ.\|g-h\|_{L^1}< \epsilon. Since a step function is a linear combination of rect functions, it sufficies to do so for rect functions. I’ll walk you through the proof in problem set 1. The idea is to use convolution.

Step 2. Now we have fgL1fhL1+ghL12ϵ\|f - g\|_{L^1} \leq \|f-h\|_{L^1}+\|g-h\|_{L^1}\leq 2\epsilon by triangle inequality. We are done.

Note that the approximation lemma holds for the more general Lebesgue integrals, the proof is actually similar. The integrals are designed to be approximated by simple functions.