Week 13: Parabolic equations

We now turn our attention to parabolic equations. To this end, suppose LL and ff are as in Week 7 (in divergence or nondivergence form). A linear second-order parabolic equation is an equation of the form

tu+Lu=f,\partial_{t}u + Lu = f,

where u=u(x,t)u=u(x,t) is a function u:Ω×IRu:\Omega\times I\rightarrow\mathbb{R}, IRI\subset\mathbb{R} an interval. Here we allow the coefficients aija_{ij}, bib^{i} and cc of LL to depend on tt as well, i.e. they are also functions on Ω×I\Omega\times I.

Motivation

Parabolic equations arise naturally in the study of heat conduction. For example, the work of Fourier implies that in an isotropic medium ΩRn\Omega\subset\mathbb{R}^{n}, given an initial temperature distribution u0:ΩRu_{0}:\Omega\rightarrow\mathbb{R}, the temperature u(x,t)u(x,t) at xΩx\in\Omega and time t>0t>0 should solve the initial-value problem

tuΔu=0limt0u(,0)=u0;\begin{aligned}\partial_{t}u-\Delta u &= 0\\ \lim_{t\searrow 0}u(\cdot,0)&=u_{0}\end{aligned};

as we will see, the solution to this problem is uniquely determined by u0u_{0} and appropriate boundary values of uu, which corresponds to the temperature distribution outside Ω\Omega. As tt\rightarrow\infty, one would expect the temperature of Ω\Omega to reach an equilibrium so that formally, the limit u:=limtu(,t)u_{\infty}:=\lim_{t\rightarrow\infty}u(\cdot,t) should solve the equation Δu=0-\Delta u_{\infty}=0. This suggests that a study of parabolic equations could lead to insight into elliptic equations and be another route to proving existence of solutions.

Another motivation for parabolic equations is the variational aspect of elliptic problems (see Week 7). Before describing this connection, consider the classical problem of finding the minima of some fC(Rn)f\in C^{\infty}(\mathbb{R}^{n}). As we know, if x0x_{0} is a minimiser of ff, then Df(x0)=0\D f(x_{0})=0. One way of finding x0x_{0} is the method of steepest descent (or negative gradient method), which essentially boils down to considering the system of ODE

x˙(t)=Df(x(t)), t>0x(0)=given\begin{aligned}\dot{x}(t)&=-\D f(x(t)),\ t>0\\x(0)&=\textup{given} \end{aligned}

and analysing the limit limtx(t)\lim_{t\rightarrow\infty}x(t), which ought to lead to a minimiser x0x_{0}, if one exists (and if x(t)x(t) is defined for all tt!). That this is promising is suggested by the fact that

ddtf(x(t))=Df2(x,t)0\frac{d}{dt}f(x(t))=-|\D f|^{2}(x,t)\leq 0

so that fxf\circ x is non-increasing. Note also that the solutions x0x_{0} to Df(x0)=0\D f(x_{0})=0 are equilibrium points of the above ODE system.

Going back to the variational formulation of the equation Lu=0Lu=0 in the case where bi0b^{i}\equiv 0 and uu vanishes on Ω\partial\Omega, recall that solutions to this equation arise as critical points of the energy

E(u)=12Ωi,jaijiuju+cu2E(u)=\frac{1}{2}\int_{\Omega}\sum_{i,j}a_{ij}\partial_{i}u\cdot \partial_{j}u + cu^{2}

in the sense that the `derivative of EE at uu in direction φC0(Ω)\varphi\in C_{0}^{\infty}(\Omega)' vanishes for all φ\varphi, which we write formally as:

<E(u),φ>:=ddtt=0E(u+tφ)=<Lu,φ>=0,\left<\nabla E(u),\varphi\right>:=\left.\frac{d}{dt}\right|_{t=0}E(u+t\varphi) = \left<Lu,\varphi\right>=0,

these inner products being the L2(Ω)L^{2}(\Omega) inner product. In analogy with the negative gradient method, we now look for appropriately regular u:Ω×[0,[Ru:\Omega\times\left[0,\infty\right[\rightarrow \mathbb{R} vanishing on Ω×[0,[\partial\Omega\times\left[0,\infty\right[ solving the initial-value problem

tu(x,t)=E(u(x,t)), (x,t)Ω×]0,[u(,0)=given.\begin{aligned}\partial_{t}u(x,t)&=-\nabla E(u(x,t)),\ (x,t)\in\Omega\times\left]0,\infty\right[\\u(\cdot,0)&=\textup{given}\end{aligned}.

As with the negative gradient method, we compute that for such a function uu,

ddtE(u)=ddt(12BL(u,u))=<Lu,tu>=tu20\frac{d}{dt}E(u)=\frac{d}{dt}\left(\frac{1}{2}B_{L}(u,u) \right) = \left<Lu,\partial_{t}u\right> = -|\partial_{t}u|^{2}\leq 0

so that uu indeed tends to decrease the energy associated to the problem. Here we again expect that as tt\rightarrow\infty, u(,t)u(\cdot,t) should appropriately tend to a solution to Lu=0Lu=0.

Weak maximum principle, comparison principle and uniqueness

In this section we consider classical solutions uC2,1(Ω×]0,T[)C0(Ω×[0,T])u\in C^{2,1}(\Omega\times\left]0,T\right[)\cap C^{0}(\overline\Omega\times [0,T]) to the Dirichlet initial-boundary value problem

tu+Lu=0 on Ω×]0,T[u(,0)=u0u(x,t)=g(x) for (x,t)Ω×[0,T[.(*)\begin{aligned}\partial_{t}u +L u&=0\ \textup{on }\Omega\times\left]0,T\right[\\ u(\cdot,0)&=u_{0}\\ u(x,t)&=g(x)\ \textup{for }(x,t)\in\partial\Omega\times\left[0,T\right[. \end{aligned}\tag{*}

for given u0C0(Ω)u_{0}\in C^{0}(\overline\Omega) and gC0(Ω×[0,T])g\in C^{0}(\partial\Omega\times\left[0,T\right]), where we shall assume that LL is an elliptic operator in nondivergence form:

Lu=i,j=1naijiju+i=1nbiiu+cuLu=\sum_{i,j=1}^{n}a_{ij}\partial_{i}\partial_{j}u + \sum_{i=1}^{n}b^{i}\partial_{i}u+cu

with aij,bi,cC0(Ω×[0,T])a_{ij},b^{i},c\in C^{0}(\overline{\Omega}\times[0,T]) for all i,j{1,,n}i,j\in\{1,\dots,n\}. Let QT:=Ω×]0,T[Q_{T}:=\Omega\times\left]0,T\right[ and define the parabolic boundary of QTQ_{T} as

QT:=(Ω×[0,T])(Ω×{0}).\partial'Q_{T}:=(\partial\Omega\times\left[0,T\right])\cup(\Omega\times\{0\}).

As before, we say that uu is a subsolution (or supersolution) to the problem (*) if the first equation is replaced with the inequality tu+Lu0\partial_{t}u+Lu\leq 0 (resp. tu+Lu0\partial_{t}u + Lu\geq 0).

Theorem 13.1 (weak maximum principle for c0c\equiv 0). Suppose uC2,1(Ω×]0,T[)C0(Ω×[0,T])u\in C^{2,1}(\Omega\times\left]0,T\right[)\cap C^{0}(\overline{\Omega}\times\left[0,T\right]) is a subsolution to the initial-boundary value problem (*) with c0c\equiv 0. Then

maxQTu=maxQTu=max{maxg,maxu0}.\max_{\overline{Q}_{T}}u = \max_{\partial'Q_{T}}u = \max\{\max g,\max u_{0}\}.

If instead uu is a supersolution to (*), then

minQTu=minQTu=min{ming,minu0}.\min_{\overline{Q}_{T}}u = \min_{\partial'Q_{T}}u=\min\{\min g,\min u_{0}\}.

Proof. We first consider the case where uu is a subsolution and tu+Lu<0\partial_{t}u+Lu<0.

Now suppose tu+Lu0\partial_{t}u+Lu\leq 0 and consider v(x,t):=u(x,t)εtv(x,t):=u(x,t) - \eps t for fixed ε>0\eps>0. We immediately see that

tv+Lv=ε+tu+Luε<0\partial_{t}v + Lv = -\eps + \partial_{t}u + Lu \leq -\eps < 0

so that we fall into the preceding case, whence for all (x,t)QT(x,t)\in Q_{T},

u(x,t)εtmaxQTv=maxQTv=max(x,t)QT(u(x,t)εt)maxQTu.u(x,t) - \eps t \leq \max_{\overline{Q}_{T}}v = \max_{\partial'Q_{T}}v = \max_{(x,t)\in\partial'Q_{T}}(u(x,t) - \eps t)\leq \max_{\partial'Q_{T}}u.

Taking the limit ε0\eps\searrow 0, we arrive at u(x,t)maxQTuu(x,t)\leq \max_{\partial'Q_{T}}u, i.e. supQTmaxQTu\sup_{Q_{T}}\leq \max_{\partial'Q_{T}}u, whence we are done. For the supersolution case, replace uu with u-u.

Remark 13.2. It follows immediately using the same techniques that if uu solves (*) with T=T=\infty, then

supQu=supQu\sup_{Q_{\infty}}u = \sup_{\partial'Q_{\infty}}u

and

infQu=infQu.\inf_{Q_{\infty}}u = \inf_{\partial'Q_{\infty}}u.

These expressions are of course only finite if gg is appropriately bounded.

Just as with elliptic equations, we also have a maximum principle in the case where c0c\geq 0.

Theorem 13.3 (weak maximum principle). Suppose uC2,1(Ω×]0,T[)C0(Ω×[0,T])u\in C^{2,1}(\Omega\times\left]0,T\right[)\cap C^{0}(\overline{\Omega}\times\left[0,T\right]) is a subsolution to the initial-boundary value problem (*) with c0c\geq 0. Then

maxQTumaxQTu+.\max_{\overline{Q}_{T}}u \leq \max_{\partial'Q_{T}}u^{+}.

If instead uu is a supersolution to (*), then

minQTumaxQTu.\min_{\overline{Q}_{T}}u \geq -\max_{\partial'Q_{T}}u^{-}.

In particular, if uu solves (*), then maxQTu=maxQTu\max_{Q_{T}}|u| = \max_{\partial'Q_{T}}|u|.

Proof. Suppose uu is a subsolution. As in the case where c0c\equiv 0, we first assume that tu+Lu<0\partial_{t}u + Lu<0, and we may as well suppose uC2,1(Ω×[0,T])u\in C^{2,1}(\Omega\times[0,T]). If maxu0\max u\leq 0, then we are clearly done, so suppose maxu=u(x0,t0)>0\max u = u(x_{0},t_{0})>0. If this point is in the interior of QTQ_{T} or in Ω×{T}\Omega\times \{T\}, we deduce as before that tu(x0,t0)0\partial_{t}u(x_{0},t_{0})\geq 0, Du(x0,t0)=0\D u(x_{0},t_{0})=0 and D2u(x0,t0)\D^{2}u(x_{0},t_{0}) is nonpositive definite so that

(tu+Lu)(x0,t0)=tu(x0,t0)i,j=1naij(x0,t0)iju(x0,t0)+i=1nbi(x0,t0)iu(x0,t0)+c(x0,t0)u(x0,t0)0,(\partial_{t} u + Lu)(x_{0},t_{0})=\partial_{t}u(x_{0},t_{0}) - \sum_{i,j=1}^{n}a_{ij}(x_{0},t_{0})\partial_{i}\partial_{j}u(x_{0},t_{0}) + \sum_{i=1}^{n}b^{i}(x_{0},t_{0})\partial_{i}u(x_{0},t_{0}) + c(x_{0},t_{0})u(x_{0},t_{0}) \geq 0,

which yields a contradiction. In the more general case, we consider v(x,t):=u(x,t)εtv(x,t):= u(x,t) -\eps t and note that if uu has a positive maximum, then so does vv for sufficiently small ε\eps. This then establishes the claim after taking ε0\eps\searrow 0 as in the case where c0c\equiv 0. The supersolution case follows from replacing uu with u-u. Finally, if uu is a solution, then it is both a subsolution and supersolution so that the inequality minQTumaxQTu\min_{Q_{T}}|u|\leq \max_{\partial'Q_{T}}|u| (and thus the equality) immediately follows.

Remark 13.4. In contrast to the elliptic case, maximum principle-type theorems may still be established even if cc is not necessarily nonnegative. For instance, if uu is a supersolution to (*) with u0,g0u_{0},g\geq 0, then the function v(x,t)=u(x,t)esupctv(x,t)=u(x,t)e^{\sup c \cdot t} satisfies

tv+(Lc)v=esupct(tu+Lu)+(supcc)esupctu0,\partial_{t}v + (L - c)v = e^{\sup c \cdot t}\left(\partial_{t}u + Lu\right) + (\sup c - c )\cdot e^{\sup|c|\cdot t}u \geq 0,

which allows us to conclude using Theorem 13.3 that v0v\geq 0, which in turn implies that u0u\geq 0. If instead uu is a subsolution and u0,g0u_{0},g\geq 0, we deduce similarly that u0u\leq 0. In this case boundedness of cc is crucial and follows as a result of our assumptions.

Corollary 13.5 (comparison principle). Suppose that uiC2,1(Ω×]0,T[)C0(Ω×[0,T])u^{i}\in C^{2,1}(\Omega\times\left]0,T\right[)\cap C^{0}(\overline{\Omega}\times\left[0,T\right]) is a solution to the initial-boundary value problem

tui+Lui=fi on Ω×]0,T[ui(,0)=u0iui(x,t)=gi(x) for (x,t)Ω×[0,T](**)\begin{aligned}\partial_{t}u^{i} +L u^{i}&=f^{i}\ \textup{on }\Omega\times\left]0,T\right[\\ u^{i}(\cdot,0)&=u_{0}^{i}\\ u^{i}(x,t)&=g^{i}(x)\ \textup{for }(x,t)\in\partial\Omega\times\left[0,T\right] \end{aligned}\tag{**}

fiC2,1(Ω×]0,T[)f^{i}\in C^{2,1}(\Omega\times\left]0,T\right[), u0iC0(Ω)u_{0}^{i}\in C^{0}(\overline{\Omega}) and giC0(Ω×[0,T])g^{i}\in C^{0}(\partial\Omega\times\left[0,T\right]) (i{1,2}i\in\{1,2\}). Suppose furthermore that the inequalities

f1f2u01u02g1g2\begin{aligned}f^{1}&\leq f^{2}\\ u_{0}^{1}&\leq u_{0}^{2}\\g^{1}&\leq g^{2} \end{aligned}

hold. Under these conditions, u1u2u^{1}\leq u_{2} on Ω×[0,T]\overline\Omega\times\left[0,T\right].

Proof. Set u:=u1u2u:=u^{1}-u^{2}. It follows immediately that uu is a subsolution to (*) with initial-boundary data u00u_{0}\leq 0 and g0g\leq 0, whence the maximum principle (cf. Remark 13.4) implies that

u1u2=umax{maxu0,maxg}0u^{1}-u^{2} = u \leq \max\{\max u_{0},\max g\}\leq 0

on QTQ_{T}, which establishes the claim.

Remark 13.6. The comparison principle implies in particular that a function uC2,1(QT)C0(QT)u\in C^{2,1}(Q_{T})\cap C^{0}(\overline{Q_{T}}) is uniquely determined by the values of tu+Lu\partial_{t}u+Lu and uQT\left.u\right|_{\partial'Q_{T}}.

Harnack inequality

Just as with elliptic equations, we have the following Harnack inequality.

Theorem 13.7 (parabolic Harnack inequality). Let UΩU\Subset\Omega be connected and t1,t2]0,T]t_{1},t_{2}\in\left]0,T\right] with t1<t2t_{1}<t_{2}. There exists a constant C>0C>0 depending only on UU, t1t_{1}, t2t_{2} and LL such that whenever uC2,1(QT)u\in C^{2,1}(Q_{T}) is a nonnegative solution to tu+Lu=0\partial_{t}u + Lu=0 on QTQ_{T},

supUu(,t1)CinfUu(,t2).\sup_{U}u(\cdot,t_{1})\leq C\cdot \inf_{U}u(\cdot,t_{2}).

Strong maximum principle

Suppose throughout this section that Ω\Omega is connected.

Theorem 13.8 (strong maximum principle for c0c\equiv 0). Suppose uC2,1(Ω×]0,T[)C0(Ω×[0,T])u\in C^{2,1}(\Omega\times\left]0,T\right[)\cap C^{0}(\overline{\Omega}\times\left[0,T\right]) is a subsolution to the initial-boundary value problem (*) with c0c\equiv 0 and maxQTu=u(x0,t0)\max_{\overline{Q}_{T}}u=u(x_{0},t_{0}) for some (x0,t0)QT(x_{0},t_{0})\in Q_{T}. Then uu is constant on Qt0Q_{t_{0}}.

If instead uu is a supersolution to (*), then if minQTu=u(x0,t0)\min_{\overline{Q}_{T}}u=u(x_{0},t_{0}) for some (x0,t0)QT(x_{0},t_{0})\in Q_{T}, we must have that uu is constant on Qt0Q_{t_{0}}.

Proof. As usual we suppose uu is a subsolution. Let

Ω={xΩ:t]0,t0[, u(x,t)=u(x0,t0)}\Omega'=\{x\in \Omega:\forall t\in \left]0,t_{0}\right[,\ u(x,t)=u(x_{0},t_{0}) \}

This set is clearly closed in Ω\Omega. We shall now show that it is open and nonempty. To this end, fix a open ball UΩU\Subset\Omega with x0Ux_{0}\in U and let vC2,1(RT)C0(RT)v\in C^{2,1}(R_{T})\cap C^{0}(\overline{R}_{T}) be a solution to the initial-boundary value problem

tv+Lv=0 on RTvRT=uRT,\begin{aligned}\partial_{t}v + Lv&=0\ \textup{on }R_{T}\\\left.v\right|_{\partial'R_{T}}&=\left.u\right|_{\partial'R_{T}}, \end{aligned}

where RT:=U×]0,T[R_{T}:=U\times \left]0,T\right[. By the comparison principle, we have that uvu(x0,t0)u\leq v\leq u(x_{0},t_{0}) so that v(x0,t0)=u(x0,t0)v(x_{0},t_{0})=u(x_{0},t_{0}) as well. Now, since u(x0,t0)v0u(x_{0},t_{0})-v\geq 0 also solves tv+Lv=0\partial_{t}v+Lv=0, we may employ the Harnack inequality on some VUV\Subset U connected with x0Vx_{0}\in V to deduce that for any t<t0t<t_{0},

supV(u(x0,t)v)C(t)infV(u(x0,t0)v)C(t)(u(x0,t0)v(x0,t0))=0\sup_{V}(u(x_{0},t) - v) \leq C(t)\inf_{V}(u(x_{0},t_{0}) - v) \leq C(t) (u(x_{0},t_{0})- v(x_{0},t_{0})) = 0

so that vu(x0,t0)v\geq u(x_{0},t_{0}), i.e. vu(x0,t0)v\equiv u(x_{0},t_{0}) so that uRt0u(x0,t0)\left.u\right|_{\partial'R_{t_{0}}}\equiv u(x_{0},t_{0}). Since Vx0V\ni x_{0} is otherwise arbitrary, it follows that uu(x0,t0)u\equiv u(x_{0},t_{0}) on Rt0R_{t_{0}}, i.e. UΩU\subset \Omega'. Ω\Omega' is therefore nonempty and, by the preceding argument, open in Ω\Omega. By connectedness, we must have Ω=Ω\Omega'=\Omega.

As in the elliptic case, we also obtain a strong maximum principle in the case where c0c\geq 0 under appropriate sign conditions.

Theorem 13.9 (string maximum principle for c0c\geq 0). Suppose uC2,1(Ω×]0,T[)C0(Ω×[0,T])u\in C^{2,1}(\Omega\times\left]0,T\right[)\cap C^{0}(\overline{\Omega}\times\left[0,T\right]) is a subsolution to the initial-boundary value problem (*) with c0c\geq 0 and maxQTu=u(x0,t0)0\max_{\overline{Q}_{T}}u=u(x_{0},t_{0})\geq 0 for some (x0,t0)QT(x_{0},t_{0})\in Q_{T}. Then uu is constant on Qt0Q_{t_{0}}.

If instead uu is a supersolution to (*), then if minQTu=u(x0,t0)0\min_{\overline{Q}_{T}}u=u(x_{0},t_{0})\leq 0 for some (x0,t0)QT(x_{0},t_{0})\in Q_{T}, we must have that uu is constant on Qt0Q_{t_{0}}.

Proof. Let Ω\Omega' be as in the preceding proof and again suppose uu is a subsolution. Note that if u(x0,t0)=0u(x_{0},t_{0})=0, the above proof applies mutatis mutandis, so suppose that u(x0,t0)>0u(x_{0},t_{0})>0. Again let UΩU\Subset \Omega be an open ball with x0Ux_{0}\in U and suppose vC2,1(RT)C0(RT)v\in C^{2,1}(R_{T})\cap C^{0}(\overline{R}_{T}) solve the initial-boundary balue problem

tv+(Lc)v=0 on RTvRT=u+RT.\begin{aligned}\partial_{t}v + (L-c)v&=0\ \textup{on }R_{T}\\ \left.v\right|_{\partial'R_{T}}&=\left.u^{+}\right|_{\partial'R_{T}}.\end{aligned}

By the weak maximum principle, we see that 0vu(x0,t0)0\leq v\leq u(x_{0},t_{0}). Moreover, on the set {u0}\{u\geq 0\}, we have that tu+(Lc)ucu0\partial_{t}u + (L-c)u\leq -cu\leq 0 so that by the comparison principle, uvu\leq v on {u0}RT\{u\geq 0\}\cap R_{T}, thus also on all of RTR_{T}, whence v(x0,t0)=u(x0,t0)v(x_{0},t_{0})=u(x_{0},t_{0}). Thus, since u(x0,t0)v0u(x_{0},t_{0})-v\geq 0 solves the equation tv+(Lc)u=0\partial_{t}v + (L-c)u=0, we again employ the Harnack inequality on VUV\Subset U connected with x0Vx_{0}\in V to obtain that for any t<t0t<t_{0},

supV(u(x0,t)v)C(t)infV(u(x0,t0)v)C(t)(u(x0,t0)v(x0,t0))=0\sup_{V}(u(x_{0},t)- v) \leq C(t)\inf_{V}(u(x_{0},t_{0}) - v)\leq C(t)(u(x_{0},t_{0}) - v(x_{0},t_{0}))=0

so that vu(x0,t0)v\geq u(x_{0},t_{0}). Therefore, as before, we must have vu(x0,t0)v\equiv u(x_{0},t_{0}) on Rt0R_{t_{0}} so that uu(x0,t0)>0u\equiv u(x_{0},t_{0})>0 on {u0}(V×]0,t0[)\{u\geq 0\}\cap \partial'(V \times ]0,t_{0}[), whence we must have uu(x0,t0)u\equiv u(x_{0},t_{0}) on all of (V×]0,t0[\partial'(V\times \left]0,t_{0}\right[. We therefore deduce as before that uRt0u(x0,t0)\left.u\right|_{R_{t_{0}}}\equiv u(x_{0},t_{0}) i.e. UΩU\subset \Omega' so that arguing as before, Ω=Ω\Omega'=\Omega.