Week 11: Maximum principles and the eigenvalue problem

We now turn our attention to classical solutions to second-order elliptic PDE and employ more classical techniques to glean information on their behaviour. Throughout this lecture, we shall suppose ΩRn\Omega\subset\mathbb{R}^{n} is open and bounded.

Write LL for the non-divergence form operator

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

where aija_{ij}, bib^{i} and cc are continuous functions on Ω\overline\Omega, (aij(x))i,j=1n(a_{ij}(x))_{i,j=1}^{n} is a symmetric matrix for each xΩx\in\Omega and for some constant λ0>0\lambda_{0}>0 the bound

i,j=1naij(x)vivjλ0v2\sum_{i,j=1}^{n}a_{ij}(x)v^{i}v^{j}\geq \lambda_{0}|v|^{2}

holds for all xΩx\in\Omega and vRnv\in\mathbb{R}^{n}.

If a function uC2(Ω)u\in C^{2}(\Omega) satisfies the inequality Lu0Lu\leq 0 on Ω\Omega, we call uu a subsolution to the equation Lu=0Lu=0, and if Lu0Lu\geq 0, we say that uu is a supersolution.

The weak maximum principle

It turns out that in the absence of the zeroth order term of LL, the pointwise behaviour of subsolutions and supersolutions are in a sense controlled in terms of their boundary data. We shall require some auxiliary results from calculus and linear algebra to establish this.

Lemma 11.1. If uC2(Ω)u\in C^{2}(\Omega) attains its maximum in Ω\Omega, i.e. if there is an x0Ωx_{0}\in\Omega such that supΩu=u(x0)\sup_{\Omega}u = u(x_{0}), then Du(x0)=0\D u(x_{0})=0 and the Hessian D2u(x0)\D^{2}u(x_{0}) is nonpositive definite.

Proof. Fix vRnv\in\mathbb{R}^{n} and consider tu(x0+tv)t\mapsto u(x_{0}+tv) for tt sufficiently close to 00. By classical calculus results, we must have that i=1nviiu(x0)=ddtt=0u(x0+tv)=0\sum_{i=1}^{n}v^{i}\partial_{i}u(x_{0})=\left.\frac{d}{d t}\right|_{t=0}u(x_{0}+tv)=0 and i,j=1nviiju(x0)vj=d2dt2t=0u(x0+tv)=0\sum_{i,j=1}^{n}v^{i}\partial_{i}\partial_{j}u(x_{0})v^{j}=\left.\frac{d^{2}}{dt^{2}}\right|_{t=0}u(x_{0}+tv)=0. Since vv is arbitrary, this establishes the claim.

Lemma 11.2. Suppose AA and BB are n×nn\times n non-negative definite matrices and AA is symmetric. Then

tr(ATB)=i,j=1naijbij0.\textup{tr}(A^{T}B) = \sum_{i,j=1}^{n}a_{ij}b_{ij}\geq 0.

Proof. By classical linear algebra, there exists an orthonormal basis {vi}i=1n\{v_{i}\}_{i=1}^{n} for Rn\mathbb{R}^{n} and non-negative numbers {λi}i=1n\{\lambda_{i}\}_{i=1}^{n} (not all distinct) such that Avi=λiviAv_{i}=\lambda_{i}v_{i}. By definition of the trace,

tr(ATB)=i=1nviTATBvi=i=1nλiviTBvi0.\textup{tr}(A^{T}B) = \sum_{i=1}^{n}v_{i}^{T}A^{T}Bv_{i} = \sum_{i=1}^{n}\lambda_{i}v_{i}^{T}Bv_{i}\geq 0.

Theorem 11.3 (weak maximum principle). Suppose uC2(Ω)C0(Ω)u\in C^{2}(\Omega)\cap C^{0}(\overline{\Omega}). If uu is a subsolution to the equation Lu=0Lu=0 with c0c\equiv 0, then it must attain its maximum on Ω\partial\Omega, i.e.

maxΩu=maxΩu.\max_{\overline\Omega}u = \max_{\partial\Omega}u.

If instead uu is a supersolution to the equation Lu=0Lu=0 with c0c\equiv 0, then it must attention its minimum on Ω\partial\Omega, i.e.

minΩu=minΩu.\min_{\overline\Omega}u = \min_{\partial\Omega}u.

Proof. First suppose that uu is a strict subsolution in the sense that Lu<0Lu<0 on Ω\Omega and assume there exists an x0Ωx_{0}\in\Omega such that maxΩu=u(x0)\max_{\Omega}u=u(x_{0}). By the preceding lemmas, we compute that

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

which contradicts the assumption that Lu<0Lu<0 on Ω\Omega so that we must have x0Ωx_{0}\in\partial\Omega. More generally, suppose Lu0Lu\leq 0 and define v:ΩRv:\Omega\rightarrow\mathbb{R} by v(x)=u(x)+εekx1v(x)=u(x) + \eps e^{kx^{1}} for fixed ε>0\eps>0 and kRk\in\mathbb{R} to be determined. We compute that

Lv(x)=Lu(x)+εekx1(k2a11(x)+kb1(x))εekx1(k2λ0+ksupb1)<0Lv(x) = Lu(x) + \eps e^{kx^{1}}\left(- k^{2}a_{11}(x) + k b^{1}(x) \right)\leq \eps e^{kx^{1}}\left(-k^{2}\lambda_{0} + k\sup b^{1} \right)<0

provided k>max{0,b1λ0}k>\max\{0,\frac{b^{1}}{\lambda_{0}}\}. Therefore, we see that for such kk,

maxxΩu(x)maxxΩ(u(x)+εekx1)maxxΩ(u(x)+εekx1)maxΩu+εmaxxΩekx1ε0maxΩu.\max_{x\in\overline\Omega}u(x) \leq \max_{x\in\overline\Omega}\left(u(x) + \eps e^{kx^{1}}\right)\leq \max_{x\in\partial\Omega}\left(u(x) + \eps e^{kx^{1}} \right)\leq \max_{\partial\Omega} u + \eps \max_{x\in\partial\Omega}e^{kx^{1}}\xrightarrow{\eps\searrow 0} \max_{\partial\Omega}u.

On the other hand, we clearly have maxΩumaxΩu\max_{\overline{\Omega}}u \geq \max_{\partial\Omega} u. This establishes the claim for subsolutions. If uu is a supersolution, then clearly u-u is a subsolution so that

minΩu=maxΩ(u)=maxΩ(u)=minΩu.\min_{\overline{\Omega}}u = -\max_{\overline{\Omega}}(-u) = -\max_{\partial\Omega}(-u) = \min_{\partial\Omega}u.

Remark 11.4. We deduce that if Lu=0Lu=0 on Ω\Omega, then uu attains its minimum and maximum on Ω\partial\Omega.

In the case where c0c\geq 0, we can apply the weak maximum principle to extract the following maximum principle.

Corollary 11.5. Suppose uC2(Ω)C0(Ω)u\in C^{2}(\Omega)\cap C^{0}(\overline{\Omega}). If uu is a subsolution to the equation Lu=0Lu=0 with c0c\geq 0, then we have the inequality

maxΩumaxΩu+,\max_{\overline\Omega}u \leq \max_{\partial\Omega}u^{+},

where u+(x):=max{u(x),0)}u^{+}(x):=\max\{u(x),0)\}. If instead uu is a supersolution to the equation Lu=0Lu=0 with c0c\geq 0, then it satisfies the inequality

minΩumaxΩu,\min_{\overline\Omega}u \geq -\max_{\partial\Omega}u^{-},

where u(x):=min{u(x),0}u^{-}(x):=-\min\{u(x),0\}.

Proof. First note that we have the inequality uu+u\leq u^{+} on Ω\overline{\Omega} so that

maxΩumaxΩu+.\max_{\overline\Omega}u\leq \max_{\overline\Omega}u^{+}.

Let U:={xΩ:u+>0}U:=\{x\in\Omega: u^{+}>0\}. If UU is empty, then u<0u<0 and we are done. Otherwise, maxΩu+=maxUu+\max_{\overline\Omega}u^{+}= \max_{\overline{U}}u^{+}, and since on UU,

i,j=1naijiju+i=1nbiiucu0,-\sum_{i,j=1}^{n}a_{ij}\cdot \partial_{i}\partial_{j}u + \sum_{i=1}^{n} b^{i}\cdot \partial_{i}u \leq -cu\leq 0,

the weak maximum principle implies that

maxΩumaxUu+=maxUu+.\max_{\overline\Omega}u \leq \max_{\overline{U}}u^{+} = \max_{\partial U} u^{+}.

Since U=(UΩ){xΩ:u(x)=0}\partial U = (\overline{U}\cap\partial\Omega)\cap \{x\in\Omega:u(x)=0\}, we see that maxUu+=maxΩu+\max_{\partial U}u^{+}=\max_{\partial\Omega}u^{+}, establishing the claim for subsolutions. As for supersolutions, we again replace uu with u-u and apply the above argument again to obtain

minΩu=maxΩ(u)maxΩ(u)+=minΩuminΩuminΩu.-\min_{\overline\Omega}u = \max_{\overline\Omega}(-u) \leq \max_{\partial\Omega} (-u)^{+} = \min_{\partial\Omega} u^{-}\Leftrightarrow \min_{\overline\Omega}u\geq -\min_{\partial\Omega}u^{-}.

The strong maximum principle

In the case where c0c\equiv 0, the weak maximum principle tells us that a (sub)solution to Lu=0Lu=0 on Ω\Omega attains its maximum on Ω\partial\Omega. We now show that if the maximum is actually attained in the interior of Ω\Omega, then uu must actually be constant. We first prove the following technical lemma.

Lemma 11.6 (Hopf). Suppose uC2(Ω)C1(Ω)u\in C^{2}(\Omega)\cap C^{1}(\overline{\Omega}) is a subsolution to the equation Lu=0Lu=0 on Ω\Omega and there exists an x0Ωx_{0}\in\partial\Omega such that u(x0)>u(x)u(x_{0})>u(x) for all xΩx\in \Omega. If Ω\Omega satisfies the interior ball condition at x0x_{0}, i.e. there exists an open ball BΩB\subset\Omega with x0Bx_{0}\in\partial B, and if either c0c\equiv 0 or c0c\geq 0 and u(x0)0u(x_{0})\geq 0, then uν(x0)>0\frac{\partial u}{\partial \nu}(x_{0})>0, where ν\nu is the outer unit normal to BB at x0x_{0}.

Proof. We may assume without loss of generality that we are dealing with the second case (otherwise replace uu with uu(x0)u-u(x_{0})). Moreover, we may suppose that B=B(0,r)B=B(0,r) for some r>0r>0. Now, define v:B(0,r)Rv:\overline{B(0,r)}\rightarrow\mathbb{R} by

v(x)=ekx2ekr2.v(x)=e^{-k|x|^{2}} - e^{-k r^{2}}.

Note that vB(0,r)0\left.v\right|_{\partial B(0,r)}\equiv 0. Moreover, we compute that iv(x)=2kxiekx2\partial_{i}v(x)=-2kx^{i} e^{-k|x|^{2}} and ijv(x)=2kδijekx2+4k2xixjekx2\partial_{i}\partial_{j}v(x)=-2k\delta_{ij}e^{-k|x|^{2}}+4k^{2}x^{i}x^{j}e^{-k|x|^{2}} so that

Lv(x)=ekx2(2k(tr (aij(x))i,j=1ni=1nbi(x)xi)4k2i,j=1naij(x)xixj+c(x))c(x)ekr2ekx2(4λ0x2k2+2k(tr (aij(x))i,j=1ni=1nbi(x)xi+c(x))).Lv(x)=e^{-k|x|^{2}}\left(2k\cdot\left(\textup{tr}\ (a_{ij}(x))_{i,j=1}^{n} - \sum_{i=1}^{n}b^{i}(x)x^{i}\right) - 4k^{2}\sum_{i,j=1}^{n}a_{ij}(x)x^{i}x^{j} + c(x) \right) - c(x) e^{-kr^{2}}\leq e^{-k|x|^{2}}\left(-4\lambda_{0}|x|^{2}k^{2} + 2k\cdot \left(\textup{tr}\ (a_{ij}(x))_{i,j=1}^{n} - \sum_{i=1}^{n}b^{i}(x)x^{i} + c(x)\right) \right).

In particular, if xB(0,r)\B(0,r2)x\in B(0,r)\backslash \overline{B(0,\frac{r}{2})}, we have that

Lv(x)ekx2(λ0r2k2+2k(suptr (aij)+supbr+supc))0Lv(x)\leq e^{-k|x|^{2}}\left(-\lambda_{0}r^{2}k^{2} + 2k\cdot\left(\sup|\textup{tr}\ (a_{ij})| + \sup|b|\cdot r + \sup |c| \right) \right)\leq 0

for sufficiently large kk. Moreover, setting w(x)=u(x)+εv(x)u(x0)w(x)=u(x)+ \eps v(x) - u(x_{0}) for sufficiently small ε>0\eps>0 such that u(x0)u(x)+εv(x)u(x_{0})\geq u(x) + \eps v(x) on B(0,r2)\partial B(0,\frac{r}{2}), since

Lw=Lu+εLvcu(x0)0,Lw = Lu + \eps Lv - cu(x_{0}) \leq 0,

we may apply the weak maximum principle to ww to conclude that for all xB(0,r)\B(0,r2)x\in \overline{B(0,r)}\backslash B(0,\frac{r}{2}),

w(x)0.w(x)\leq 0.

In particular, we see that

wν(x0)=limh0w(x0hν)w(x0)h=limh0w(x0hν)h0uν(x0)εvν(x0)=2εkrekr2>0.\frac{\partial w}{\partial \nu}(x_{0})=\lim_{h\searrow 0} \frac{w(x_{0}-h\nu)-w(x_{0}) }{-h} = \lim_{h\searrow 0}\frac{-w(x_{0}-h\nu) }{h}\geq 0\Leftrightarrow \frac{\partial u}{\partial \nu}(x_{0})\geq -\eps\frac{\partial v}{\partial \nu}(x_{0})=2\eps k r\cdot e^{-k r^{2}}>0.

Theorem 11.7 (strong maximum principle). Suppose uC2(Ω)C0(Ω)u\in C^{2}(\Omega)\cap C^{0}(\overline\Omega), c0c\equiv 0 on Ω\Omega and Ω\Omega is connected. If uu is a subsolution to the equation Lu=0Lu=0 on Ω\Omega and maxΩu=u(x0)\max_{\overline{\Omega}}u=u(x_{0}) for some x0Ωx_{0}\in \Omega, then uu is constant. Likewise, if uu is a supersolution to the equation Lu=0Lu=0 on Ω\Omega and minΩu=u(x0)\min_{\overline{\Omega}}u=u(x_{0}) for some x0Ωx_{0}\in\Omega, then uu is constant.

Proof. Suppose uu is a subsolution and let C:={xΩ:u(x)=maxΩu}C:=\{x\in\Omega:u(x)=\max_{\overline{\Omega}}u\} and U:=Ω\C={xΩ:u(x)<maxΩu}U:=\Omega\backslash C=\{x\in\Omega:u(x)<\max_{\overline{\Omega}}u\}. Suppose UU is nonempty, fix yUy\in U such that dist(y,C)<dist(y,Ω)\dist(y,C)<\dist(y,\partial\Omega) and let BB be the largest ball centred at yy such that BUB\subset U so that BC\partial B\cap C is nonempty. Let y0BCy_{0}\in \partial B\cap C. By construction, UU satisfies the interior ball condition at y0y_{0} so that we may apply the Hopf lemma to conclude that uν(y0)>0\frac{\partial u}{\partial \nu}(y_{0})>0, but y0y_{0} is a local maximum so that Du(y0)=0\D u(y_{0})=0, which is a contradiction.

Similarly, we have the following result in the case where c0c\geq 0. The proof is analogous to that of Theorem 11.7, using the latter part of the Hopf lemma.

Theorem 11.8 (strong maximum principle). Suppose uC2(Ω)C0(Ω)u\in C^{2}(\Omega)\cap C^{0}(\overline\Omega), c0c\geq 0 on Ω\Omega and Ω\Omega is connected. If uu is a subsolution to the equation Lu=0Lu=0 on Ω\Omega and maxΩu=u(x0)0\max_{\overline{\Omega}}u=u(x_{0})\geq 0 for some x0Ωx_{0}\in \Omega, then uu is constant. Likewise, if uu is a supersolution to the equation Lu=0Lu=0 on Ω\Omega and minΩu=u(x0)0\min_{\overline{\Omega}}u=u(x_{0})\leq 0 for some x0Ωx_{0}\in\Omega, then uu is constant.

Harnack inequality

In addition to the maximum principles, we have the following inequality for nonnegative solutions to Lu=0Lu=0 on Ω\Omega, which tells us that the supremum of such a solution is controlled by its infimum away from Ω\partial\Omega.

Theorem 11.9. For any connected UΩU\Subset \Omega, there exists a constant C>0C>0 depending only on UU and LL such that whenever uC2(Ω)u\in C^{2}(\Omega) is nonnegative and solves Lu=0Lu=0 on Ω\Omega,

supUuCinfUu.\sup_{U}u\leq C\inf_{U}u.