Week 12: Eigenvalues and eigenfunctions

Throughout this lecture, Ω\Omega is a connected bounded domain of class CC^{\infty}.

The eigenvalue problem

We now turn our attention to the eigenvalue problem for LL, i.e. the problem of finding λR\lambda\in\mathbb{R} and uu not identically zero solving the following boundary value problem:

Lu=λu on ΩuΩ0\begin{aligned} Lu&=\lambda u\ \textup{on }\Omega\\\left.u\right|_{\partial\Omega}&\equiv 0 \end{aligned}

By Theorem 9.3, there are at most countably many eigenvalues λ\lambda (and correponding eigenfunctions uu). We now obtain more refined information on the solutions to this eigenvalue problem.

Assume LL is in divergence form with the same setup as in Week 8 and that aijC(Ω)a_{ij}\in C^{\infty}(\overline{\Omega}). We shall focus on the case where

Lu=i,j=1ni(aijju),Lu= -\sum_{i,j=1}^{n}\partial_{i}(a_{ij}\cdot\partial_{j}u),

in which case LuLuL^{\ast}u\equiv Lu, since BL(u,v)=Ωi,jaijiujv=BL(u,v)B_{L}(u,v)=\int_{\Omega}\sum_{i,j}a_{ij}\partial_{i}u\partial_{j}v=B_{L^{\ast}}(u,v) for any u,vH01,2(Ω)u,v\in H_{0}^{1,2}(\Omega); we say that LL is symmetric. We shall require the following fundamental result from functional analysis.

Theorem 12.1 (eigenvectors of symmetric compact operators). If XX is a separable Hilbert space (i.e. XX has a countable dense subset) and A:XXA:X\rightarrow X is a compact symmetric linear mapping (for all x,yXx,y\in X <x,Ay>=<Ax,y>\left<x,Ay\right>=\left<Ax,y\right>), then the eigenvalues {μi}i=1\{\mu{i}\}_{i=1}^{\infty} (counted with multiplicity) of AA are real and there exists an orthonormal basis {φi}i=1\{\varphi_{i}\}_{i=1}^{\infty} of XX such that for each iNi\in\mathbb{N}, Aφi=μiφiA \varphi_{i}=\mu_{i}\varphi_{i}.

Theorem 12.2 (eigenvalues of symmetric elliptic operators). If b0b\equiv 0 and c0c\equiv 0, then each eigenvalue λ\lambda of LL is real and positive. Moreover, writing {λk}k=1\{\lambda_{k}\}_{k=1}^{\infty} for the eigenvalues of LL counted with multiplicity, we have that 0<λ1λ20<\lambda_{1}\leq \lambda_{2}\leq \dots with λkk\lambda_{k}\xrightarrow{k\rightarrow\infty}\infty. Finally, there exists a sequence {wk}k=1H01,2(Ω)\{w_{k}\}_{k=1}^{\infty}\subset H_{0}^{1,2}(\Omega) with Lwk=λkwkLw_{k}=\lambda_{k}w_{k} for each kNk\in\mathbb{N} such that {wk}k=1\{w_{k}\}_{k=1}^{\infty} is an orthonormal basis of L2(Ω)L^{2}(\Omega).

Proof. First note that if Lu=λuLu=\lambda u for some λC\lambda\in \mathbb{C} and real-valued uu not equivalently zero, then

BL(u,u)=<Lu,u>=λu22λ=BL(u,u)u22=Ωi,jaijiujuu22Du22u22>0.B_{L}(u,u)=\left<Lu,u\right> = \lambda||u||_{2}^{2}\Leftrightarrow \lambda = \frac{B_{L}(u,u)}{||u||_{2}^{2}}=\frac{\int_{\Omega}\sum_{i,j}a_{ij}\cdot \partial_{i}u\partial_{j}u}{||u||_{2}^{2}}\geq \frac{||\D u||_{2}^{2}}{||u||_{2}^{2}}> 0.

Now, since BLB_{L} is coërcive, Lax-Milgram implies that there is a unique continuous linear mapping A:L2(Ω)H01,2(Ω)L2(Ω)A:L^{2}(\Omega)\rightarrow H_{0}^{1,2}(\Omega)\subset L^{2}(\Omega) such that

BL(Af,v)=<f,v>B_{L}(Af,v)=\left<f,v\right>

for all fL2(Ω)f\in L^{2}(\Omega) and vH01,2(Ω)v\in H_{0}^{1,2}(\Omega). Just as in Week 8, we deduce that A:L2(Ω)L2(Ω)A:L^{2}(\Omega)\rightarrow L^{2}(\Omega) is compact, and we see that it is symmetric as well, since whenever fL2(Ω)f\in L^{2}(\Omega) and gH01,2(Ω)g\in H_{0}^{1,2}(\Omega),

<f,Ag>=BL(Af,Ag)=BL(Ag,Af)=<g,Af>\left<f,Ag\right> = B_{L}(Af,Ag) = B_{L}(Ag,Af) = \left<g,Af\right>

so that by approximation, <f,Ag>=<Af,g>\left<f,Ag\right> = \left<Af,g\right> for all f,gL2(Ω)f,g\in L^{2}(\Omega). Now,

uH01,2(Ω), Lu=λuuL2(Ω),Au=λ1uu\in H_{0}^{1,2}(\Omega),\ Lu=\lambda u \Leftrightarrow u\in L^{2}(\Omega), Au = \lambda^{-1}u

so that as in Weeks 8 and 9, we deduce that the set of all eigenvalues of LL may be written as monotone nondecreasing sequence {λi}i=1\{\lambda_{i}\}_{i=1}^{\infty} with λii\lambda_{i}\xrightarrow{i\rightarrow\infty}\infty. By Theorem 11.10, the (orthonormal) eigenfunctions {wi}i=1\{w_{i}\}_{i=1}^{\infty} associated with {λi}i=1\{\lambda_{i}\}_{i=1}^{\infty} span L2(Ω)L^{2}(\Omega).

Remark 12.3. By higher regularity, the {uk}k=1\{u_{k}\}_{k=1}^{\infty} are smooth on Ω\Omega (up to a set of measure zero), and if Ω\partial\Omega is smooth, they are smooth on Ω\overline\Omega.

By Theorem 12.2, any uL2(Ω)u\in L^{2}(\Omega) may be written in terms of the eigenfunctions {wk}k=1\{w_{k}\}_{k=1}^{\infty} as

u=L2limNi=1Nuiwi,u=L^{2}-\lim_{N\rightarrow\infty} \sum_{i=1}^{N}u_{i}w_{i},

where ui=<wi,u>u_{i}=\left<w_{i},u\right> for each iNi\in\mathbb{N}. We now turn our attention to the convergence of this series in H01,2(Ω)H_{0}^{1,2}(\Omega) for a given uH01,2(Ω)u\in H_{0}^{1,2}(\Omega). We first note the following.

Lemma 12.4. The bilinear form BLB_{L} associated to LL defines an inner product on H01,2(Ω)H_{0}^{1,2}(\Omega). Moreover, the associated norm is equivalent to 1,2||\cdot||_{1,2}, i.e. there exist constants c0,c1>0c_{0},c_{1}>0 depending only on Ω\Omega such that for any uH01,2(Ω)u\in H_{0}^{1,2}(\Omega),

c0u1,2BL(u,u)c1u1,2.c_{0}||u||_{1,2}\leq \sqrt{B_{L}(u,u)}\leq c_{1}||u||_{1,2}.

Proof. BLB_{L} is bilinear and symmetric by assumption. Moreover, by ellipticity,

BL(u,u)λ0ΩDu2λ02min{CPoinc1,1}u1,22B_{L}(u,u)\geq \lambda_{0}\int_{\Omega}|\D u|^{2}\geq \frac{\lambda_{0}}{2}\min\{C_{\textup{Poinc}}^{-1},1\}||u||_{1,2}^{2}

so that positive-definiteness is immediate. The equivalence claim now follows from this inequality and the boundedness of (aij)(a_{ij}):

BL(u,u)sup(aij)ΩDu2sup(aij)u1,22.B_{L}(u,u)\leq\sup|(a_{ij})|\int_{\Omega}|\D u|^{2}\leq \sup|(a_{ij})|\cdot ||u||_{1,2}^{2}.

Remark 12.5. BLB_{L} is commonly known as the energetic inner product and the associated norm the energetic norm associated to LL, the motivation being the Dirichlet-type energy

E(u)=12Ωi,j=1naijiuju=12BL(u,u)E(u)=\frac{1}{2}\int_{\Omega}\sum_{i,j=1}^{n}a_{ij}\partial_{i}u\cdot\partial_{j}u=\frac{1}{2}B_{L}(u,u)

from which BLB_{L} naturally arises (cf. Lemma 7.10).

Corollary 12.6. The eigenfunctions {wk}k=1\{w_{k}\}_{k=1}^{\infty} form a basis of H01,2(Ω)H_{0}^{1,2}(\Omega) and for any uH01,2(Ω)u\in H_{0}^{1,2}(\Omega),

u=W1,2limNi=1Nuiwi,u=W^{1,2}-\lim_{N\rightarrow\infty}\sum_{i=1}^{N}u_{i}w_{i},

where ui=<wi,u>u_{i}=\left<w_{i},u\right> for each iNi\in\mathbb{N}.

Proof. We shall employ the energetic inner product. First of all, note that by definition,

BL(wi,wj)=λiΩwiwj=λiδijBL(wiλi,wjλj)=δij,B_{L}(w_{i},w_{j})=\lambda_{i}\int_{\Omega}w_{i}w_{j}=\lambda_{i}\delta_{ij}\Leftrightarrow B_{L}(\frac{w_{i}}{\sqrt{\lambda_{i}}},\frac{w_{j}}{\sqrt{\lambda_{j}}})=\delta_{ij},

i.e. the set of functions {wiλi}i=1\{\frac{w_i}{\sqrt{\lambda_{i}}}\}_{i=1}^{\infty} is BLB_{L}-O.N. To see that their span XX is all of H01,2(Ω)H_{0}^{1,2}(\Omega), note that if uH01,2(Ω)u\in H_{0}^{1,2}(\Omega), we may write

u=u0+u1u=u_{0}+u_{1}

with u0Xu_{0}\in X and u1H01,2(Ω)u_{1}\in H_{0}^{1,2}(\Omega) such that BL(u1,v)=0B_{L}(u_{1},v)=0 for all vXv\in X. However, this implies that for each iNi\in\mathbb{N},

0=BL(u1,wi)=BL(wi,u1)=λiΩwiu1Ωwiu1=00=B_{L}(u_{1},w_{i})=B_{L}(w_{i},u_{1})=\lambda_{i}\int_{\Omega}w_{i}u_{1}\Rightarrow\int_{\Omega}w_{i}u_{1}=0

so that u1=0u_{1}=0 in L2(Ω)L^{2}(\Omega), thus also in H01,2(Ω)H_{0}^{1,2}(\Omega). Therefore, uXu\in X and we may expand uu in terms of the orthonormal basis {wiλi}i=1\{\frac{w_{i}}{\sqrt{\lambda_{i}}}\}_{i=1}^{\infty} (limit in W1,2(Ω)W^{1,2}(\Omega)):

u=i=1BL(u,wiλi)wiλi=i=1BL(u,wi)λiwi=i=1<wi,u>wi.u= \sum_{i=1}^{\infty}B_{L}(u,\frac{w_{i}}{\sqrt{\lambda_{i}}})\cdot\frac{w_{i}}{\sqrt{\lambda_{i}}}=\sum_{i=1}^{\infty}\frac{B_{L}(u,w_{i})}{\lambda_{i}}\cdot w_{i}=\sum_{i=1}^{\infty}\left<w_{i},u\right>\cdot w_{i}.

Example 12.7. A ubiquitous concrete instance of the above considerations is n=1n=1, Ω=]0,1[\Omega=]0,1[ and Lu:=Δu=uLu:=-\Delta u = -u''. In this case, the eigenvalue problem boils down to the problem of solving the ODE

u(x)=λu(x), x]0,1[-u''(x)=\lambda u(x),\ x\in\left]0,1\right[

for uu and λ>0\lambda>0 subject to the boundary conditions u(0)=u(1)=0u(0)=u(1)=0. From ODE theory, we know that any C2C^{2} solution to this equation is of the form

u(x)=Acos(λx)+Bsin(λx),u(x)=A\cos(\sqrt{\lambda}x) + B \sin(\sqrt{\lambda}x),

where A,BRA,B\in\mathbb{R}. Using the boundary conditions, we obtain constraints on AA, BB and λ\lambda:

0=u(0)=A0=u(1)=Bsin(λ)\begin{aligned}&0=u(0)=A\\ &0=u(1)=B\sin(\sqrt\lambda)\end{aligned}

The last condition implies that either B=0B=0, in which case uu is the trivial solution, or sinλ=0λ=jπ\sin\sqrt\lambda=0\Leftrightarrow \sqrt{\lambda}=j\pi for jNj\in\mathbb{N}. Therefore, we see that the eigenvalues of LL are given by {λj=j2π2}j=1\{\lambda_{j}=j^{2}\pi^{2}\}_{j=1}^{\infty} with corresponding orthonormal eigenfunctions {xwj(x)=2sin(jπx)}j=1\{x\mapsto w_{j}(x)=\sqrt{2}\sin(j\pi x)\}_{j=1}^{\infty} so that any uL2(]0,1[)u\in L^{2}(\left]0,1\right[) may be expanded in the form

u(x)=j=1uj2sin(jπx);u(x)=\sum_{j=1}^{\infty}u_{j}\cdot \sqrt{2}\sin(j\pi x);

such a series is called a Fourier series.

The first eigenvalue

We now take a closer look at the smallest eigenvalue λ1>0\lambda_{1}>0 of Lu=i,j=1ni(aijju)Lu=-\sum_{i,j=1}^{n}\partial_{i}(a_{ij}\partial_{j}u), which we call the principal eigenvalue. It turns out it may be characterised variationally.

Theorem 12.8 (variational principle for principal eigenvalue). The principal eigenvalue of LL is characterised by the condition

λ1=inf{BL(u,u):uH01,2(Ω),u2=1};(*)\lambda_{1}=\inf\{B_{L}(u,u):u\in H_{0}^{1,2}(\Omega),||u||_{2}=1\};\tag{*}

moreover, the infimum is attained at a positive function w1H01,2(Ω)C(Ω)w_{1}\in H_{0}^{1,2}(\Omega)\cap C^{\infty}(\Omega) solving the equation Lw1=λ1w1Lw_{1}=\lambda_{1}w_{1} , and if uH01,2(Ω)u\in H_{0}^{1,2}(\Omega) is any weak solution to the equation Lu=λ1uLu=\lambda_{1}u, then u=kw1u=k\cdot w_{1} a.e. for some kRk\in\mathbb{R}.

Remark 12.9. This theorem states in particular that the eigenvalue λ1\lambda_{1} is simple so that the eigenvalues of LL satisfy the inequality

0<λ1<λ2λ3.0<\lambda_{1}<\lambda_{2}\leq \lambda_{3}\leq \dots.

Remark 12.10. (*) is known as Rayleigh's formula and may be written as

λ1=minuH01,2(Ω)\{0}BL(u,u)u22.\lambda_{1}=\min_{u\in H_{0}^{1,2}(\Omega)\backslash\{0\}}\frac{B_{L}(u,u)}{||u||_{2}^{2}}.

If in particular L=ΔL=-\Delta, then the lowest eigenvalue is given by

λ1=minuH01,2(Ω)\{0}ΩDu2Ωu2=CPoinc(Ω)1,\lambda_{1}=\min_{u\in H_{0}^{1,2}(\Omega)\backslash\{0\}}\frac{\int_{\Omega}|\D u|^{2}}{\int_{\Omega}|u|^{2}}=C_{\textup{Poinc}}(\Omega)^{-1},

cf. Exercise 4 on tutorial sheet 3.

Proof of Theorem 12.8. Suppose uH01,2(Ω)u\in H_{0}^{1,2}(\Omega) with u2=1||u||_{2}=1. Since by Corollary 12.6, uu may be expanded as

u=j=1<wj,u>wju=\sum_{j=1}^{\infty}\left<w_{j},u\right>w_{j}

in H01,2(Ω)H_{0}^{1,2}(\Omega), we see that

BL(u,u)=j,k=1<wj,u><wk,u>BL(wj,wk)=j=1λj<wj,u>2j=1λ1<wj,u>2=λ1u22=λ1,B_{L}(u,u) = \sum_{j,k=1}^{\infty}\left<w_{j},u\right>\cdot \left<w_{k},u\right>\cdot B_{L}(w_{j},w_{k}) = \sum_{j=1}^{\infty}\lambda_{j}\cdot |\left<w_{j},u\right>|^{2}\geq \sum_{j=1}^{\infty}\lambda_{1}\cdot |\left<w_{j},u\right>|^{2} = \lambda_{1}||u||_{2}^{2}=\lambda_{1},

with equality iff <wj,u>=0\left<w_{j},u\right>=0 for all j>1j>1, i.e. iff u=w1u=w_{1}. This establishes the first claim. For the second claim, we shall show that if uH01,2(Ω)u\in H_{0}^{1,2}(\Omega) is a nontrivial weak solution to Lu=λ1uLu=\lambda_{1}u, then it must be either positive or negative on Ω\Omega. To this end, write u+(x):=max{u(x),0}u_{+}(x):=\max\{u(x),0\} and u(x):=max{u(x),0}u_{-}(x):=\max\{-u(x),0\} so that uu may be written as

u=u+u,u=u_{+}-u_{-},

and we have that u+u=0u_{+}\cdot u_{-}=0 so that u2=u+2+u2||u||^{2}= ||u_{+}||^{2}+ ||u_{-}||^{2}. Moreover, it may be shown that u±H01,2(Ω)u_{\pm}\in H_{0}^{1,2}(\Omega) and

Du±(x)={±Du(x),x{±u0}0,x{±u0}\D u_{\pm}(x)=\begin{cases}\pm \D u(x),& x\in \{\pm u \geq 0\}\\ 0,& x\in \{\pm u\leq 0\} \end{cases}

a.e. so that iu+ju0\partial_{i}u_{+}\cdot \partial_{j}u_{-}\equiv 0 a.e., which in turn implies that BL(u+,u)=0B_{L}(u_{+},u_{-})=0 so that BL(u,u)=BL(u+,u+)+BL(u,u)B_{L}(u,u) = B_{L}(u_{+},u_{+}) + B_{L}(u_{-},u_{-}). Now, by the proof of the first part, BL(u±,u±)λ1u±22B_{L}(u_{\pm},u_{\pm})\geq \lambda_{1}||u_{\pm}||_{2}^{2}. It turns out that equality holds in both cases, for if we had a strict inequality, then

λ1u22=BL(u,u)=BL(u+,u+)+BL(u,u)>λ1(u+22+u22)=λ1u22,\lambda_{1}||u||_{2}^{2} = B_{L}(u,u) = B_{L}(u_{+},u_+) + B_{L}(u_-,u_-)> \lambda_{1}\left(||u_{+}||_{2}^{2} + ||u_{-}||_{2}^{2}\right) = \lambda_{1}||u||_{2}^{2},

which is impossible. Therefore, by the first part, u+H01,2(Ω)u_{+}\in H_{0}^{1,2}(\Omega) is a weak solution to Lu=λ1uLu = \lambda_{1}u. By regularity theory, u+C(Ω)u_{+}\in C^{\infty}(\overline\Omega), and since

Lu+=λ1u+0,Lu_{+}=\lambda_{1} u_{+}\geq 0,

u+u_{+} is a supersolution to the equation Lu=0Lu=0 so that by the strong maximum principle, since u+Ω0\left.u_{+}\right|_{\partial\Omega}\equiv 0 and u+0u_{+}\geq 0, either u+0u_{+}\equiv 0 on Ω\Omega or u+>0u_{+}>0, in which case u0u_{-}\equiv 0. A similar argument applies to uu_{-} as well, thus establishing the second claim. To establish the last claim, let u,vH01,2(Ω)u,v\in H_{0}^{1,2}(\Omega) be any nontrivial solutions to Lu=λ1uLu=\lambda_{1}u. Then by the preceding part, both Ωu\int_{\Omega}u and Ωv\int_{\Omega}v are nonzero so that we may find an αR\alpha\in \mathbb{R} such that

Ωuαv=0.\int_{\Omega} u - \alpha v=0.

But since uαvu-\alpha v is a solution to the same equation, the preceding part now implies that uαv0u-\alpha v\equiv 0.

All of the above considerations were for the symmetric elliptic operator

Lu=i,j=1ni(aijju).Lu=-\sum_{i,j=1}^{n}\partial_{i}(a_{ij}\partial_{j}u).

More generally, consider the more general nondivergence form elliptic operator

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,cC(Ω)a_{ij}, b^{i},c\in C^{\infty}(\overline\Omega) for all i,j{1,,n}i,j\in\{1,\dots,n\} and c0c\geq 0. This operator is not symmetric in general; therefore, its eigenvalues will not necessarily be real. However, we can still say something about its principal eigenvalue, which is both real and positive.

Theorem 12.11. There exists a λ1>0\lambda_{1}>0 such that