Week 2: Characterisations of Sobolev spaces

Sobolev spaces (continued)

We illustrate the potentially pathological nature of Sobolev functions with two more examples.

Example 2.1. Consider the mapping u:ΩRu:\Omega\rightarrow\mathbb{R} such that xxαx\mapsto |x|^{-\alpha}, where Ω:=B(0,1)\Omega:=B(0,1). This function is smooth on Ω\{0}\Omega\backslash\{0\} and

iu(x)=αxα+2xi.\partial_{i}u(x)=-\frac{\alpha}{|x|^{\alpha+2}}x^{i}.

A quick computation shows that for p1p\geq 1,

B(0,1)Dup=nωn01rn1(1)pαprp(α+1)dr,\int_{B(0,1)}|\D u|^{p}=n\omega_{n}\cdot \int_{0}^{1}r^{n-1}\cdot\frac{(-1)^{p}\alpha^{p}}{r^{p\cdot (\alpha + 1)}}dr,

which is finite iff n1p(α+1)>1p(α+1)<nn-1-p(\alpha+1)>-1\Leftrightarrow p\cdot(\alpha+1)<n. If this is the case, then by the usual integration by parts formula,

Ω\B(0,ε)uiφ=Ω\B(0,ε)iuφ+B(0,ε)uφxiε\int_{\Omega\backslash B(0,\eps)}u\cdot \partial_{i}\varphi = -\int_{\Omega\backslash B(0,\eps)}\partial_{i}u\cdot \varphi +\int_{\partial B(0,\eps)}u\cdot \varphi\cdot \frac{x^{i}}{\eps}

for every ε>0\eps>0 and φC0(Ω)\varphi\in C_{0}^{\infty}(\Omega). Now, using the definition of uu and crudely estimating, we see that

B(0,ε)uφxiεsupφεαB(0,ε)1=nωnεn1ε00.\left|\int_{\partial B(0,\eps)}u\cdot \varphi\cdot \frac{x^{i}}{\eps} \right|\leq \sup|\varphi| \cdot \eps^{-\alpha}\cdot \underbrace{\int_{\partial B(0,\eps)}1}_{=n\cdot\omega_{n}\cdot \eps^{n-1}}\xrightarrow{\eps\searrow 0}0.

Since u,iuLloc1u,\partial_{i}u\in \lloc^{1}, we may take limits above to obtain the relation

Ωuiφ=Ωiuφ,\int_{\Omega}u\cdot \partial_{i}\varphi = - \int_{\Omega}\partial_{i}u\cdot\varphi,

i.e. whenever p(α+1)<np\cdot (\alpha+1)<n, we have that uW1,p(Ω)u\in W^{1,p}(\Omega).

Example 2.2. Let {rk}k=1Ω:=B(0,1)\{r_{k}\}_{k=1}^{\infty}\subset \Omega:=B(0,1) be dense and define u:ΩRu:\Omega\rightarrow\mathbb{R} such that xk=12kxrkαx\mapsto \sum_{k=1}^{\infty}2^{-k}\cdot |x-r_{k}|^{-\alpha}. It may be shown that uW1,p(Ω)u\in W^{1,p}(\Omega) iff p(α+1)<nα<nppp\cdot(\alpha+1)<n\Leftrightarrow \alpha<\frac{n-p}{p}. If α>0\alpha>0 as well, then uu is not bounded on any open subset of Ω\Omega.

Sobolev spaces via smooth functions

We now turn our attention to alternative characterisations of weak derivatives and Sobolev spaces.

Derivatives in the LpL^{p} sense (p<p<\infty)

Definition 2.3. Suppose uLlocp(Ω)u\in\lloc^{p}(\Omega). We say that Dαu\D^{\alpha}u exists in the LpL^{p} sense if there is a vLlocp(Ω)v\in \lloc^{p}(\Omega) such that for every UΩU\Subset\Omega, there exists a sequence {uj}j=1Cα(Ω)\{u_{j}\}_{j=1}^{\infty}\subset C^{|\alpha|}(\Omega) such that ujjuu_{j}\xrightarrow{j\rightarrow\infty}u and Dαujjv\D^{\alpha}u_{j}\xrightarrow{j\rightarrow\infty}v in Lp(U)L^{p}(U). We then say v=Dαuv=\D^{\alpha}u in the LpL^{p} sense and call vv the α\alphath LpL^{p} derivative of uu.

Exercise 2.4. If v=Dαuv=\D^{\alpha}u in the LpL^{p} sense, then v=Dαuv=\D^{\alpha}u in the weak sense.

As it turns out, the converse is also true for u,DαuLlocp(Ω)u,\D^{\alpha}u\in\lloc^{p}(\Omega).

Lemma 2.5. If uLlocp(Ω)u\in \lloc^{p}(\Omega) and Dαu\D^{\alpha}u exists in the weak sense and is in Llocp(Ω)\lloc^{p}(\Omega), then Dαu\D^{\alpha}u is the α\alphath LpL^{p} derivative of uu.

Proof. Let UΩU\Subset\Omega be fixed. We know from Theorem 1.1 that Jεuε0uJ_{\eps}u\xrightarrow{\eps\searrow 0}u in Lp(U)L^{p}(U). We now show that for sufficiently small ε>0\eps>0, DαJεu=JεDαu\D^{\alpha}J_{\eps}u=J_{\eps}\D^{\alpha}u so that uj:=J1/juu_{j}:=J_{1/j}u is the desired sequence.

Now, by definition, for each xUx\in U, by the dominated convergence theorem,

DαJεu(x)=εnΩDxα(ρ(xzε))u(z)dz=εn(1)αΩDzα(ρ(xzε))u(z)dz.\D^{\alpha}J_{\eps}u(x)=\eps^{-n}\int_{\Omega}\D^{\alpha}_{x}\left(\rho\left(\frac{x-z}{\eps}\right)\right)\cdot u(z)dz=\eps^{-n}\cdot (-1)^{|\alpha|}\int_{\Omega}\D^{\alpha}_{z}\left(\rho\left(\frac{x-z}{\eps}\right)\right)\cdot u(z)dz.

In order to shift the derivative Dzα\D_{z}^{\alpha} onto uu, we need to verify that zρ(xzε)z\mapsto \rho(\frac{x-z}{\eps}) is compactly supported in Ω\Omega for sufficiently small ε>0\eps>0. Now, we know that supp (zρ(xzε))B(x,ε)\supp (z\mapsto \rho(\frac{x-z}{\eps}))\subset B(x,\eps). Taking ε<dist(U,Ω)\eps<\dist(U,\partial\Omega), we see that B(x,ε)ΩB(x,\eps)\Subset \Omega. Therefore, we may throw the derivative onto uu:

DαJεu(x)=εnΩρ(xzε)Dαu(z)dz=JεDαu(x).\Rightarrow \D^{\alpha}J_{\eps}u(x) = \eps^{-n}\int_{\Omega}\rho\left(\frac{x-z}{\eps}\right)\cdot \D^{\alpha}u(z)dz= J_{\eps}\D^{\alpha}u(x).

The claim now follows from our remarks above.

Remark 2.6. In analogy with Llocp\lloc^{p}, we say that uWlock,p(Ω)u\in\wloc^{k,p}(\Omega) if uWlock,p(U)u\in \wloc^{k,p}(U) for every UΩU\Subset \Omega, and given a sequence {ui}Wlock,p(Ω)\{u_{i}\}\subset \wloc^{k,p}(\Omega), we say that uiiuu_{i}\xrightarrow{i\rightarrow\infty}u in Wlock,p(Ω)\wloc^{k,p}(\Omega) if uiiuu_{i}\xrightarrow{i\rightarrow\infty}u in Wk,p(U)W^{k,p}(U) for every UΩU\Subset \Omega. In particular, the above lemma tells us that for every uWlock,p(Ω)u\in \wloc^{k,p}(\Omega), we may find a sequence {ui:=J1/iu}i=1C(Ω)\{u_{i}:=J_{1/i}u\}_{i=1}^{\infty}\subset C^{\infty}(\Omega) converging to uu in Wlock,p(Ω)\wloc^{k,p}(\Omega).

The spaces Hk,pH^{k,p} and H0k,p(Ω)H_{0}^{k,p}(\Omega)

In light of the previous section, we know that we may approximate Sobolev functions locally by means of smooth functions. Recall from Theorem 1.1 that for p<p<\infty, Lp(Ω)=C0(Ω)L^{p}(\Omega)=\overline{\cs(\Omega)}. Motivated by this, we set

Ck,p(Ω):={fCk(Ω):fk,p<}C^{k,p}(\Omega):=\{f\in C^{k}(\Omega): ||f||_{k,p}<\infty\}

and introduce the subspaces

Hk,p(Ω):=Ck,p(Ω)H0k,p(Ω):=C0(Ω) \begin{aligned} H^{k,p}(\Omega)&:=\overline{C^{k,p}(\Omega)}\\ H_{0}^{k,p}(\Omega)&:=\overline{C_{0}^{\infty}(\Omega)} \end{aligned}

of Wk,p(Ω)W^{k,p}(\Omega). Note that Ck,p(Ω)C^{k,p}(\Omega) is not closed in Wk,p(Ω)W^{k,p}(\Omega).
[[Counterexample: Ω=]1,1[\Omega=]-1,1[, k=1k=1, p=1p=1, fn(x)=x2+1nf_{n}(x)=\sqrt{x^{2}+\frac{1}{n}}]]

It is clear that H0k,p(Ω)Hk,p(Ω)Wk,p(Ω)H_{0}^{k,p}(\Omega)\subset H^{k,p}(\Omega)\subset W^{k,p}(\Omega). We will show that the latter inclusion is actually an equality. We first recall the following technical lemma.

Lemma 2.7. Let {Uα}\{U_{\alpha}\} be an open cover of a set ARnA\subset\mathbb{R}^{n}. There exists a countable family of smooth functions {ψi:Rn[0,[}\{\psi_{i}:\mathbb{R}^{n}\rightarrow[0,\infty[ \} such that

If the cover {Uα}\{U_{\alpha}\} is countable and locally finite, then we may find a collection of functions {ψα}\{\psi_{\alpha}\} with the same index set as the cover satisfying the above conditions. In either case, we call the collection {ψi}\{\psi_{i}\} a partition of unity.

Theorem 2.8. If p<p<\infty, then Hk,p(Ω)=Wk,p(Ω)H^{k,p}(\Omega)=W^{k,p}(\Omega).

Proof. We will show that any uWk,p(Ω)u\in W^{k,p}(\Omega) may be approximated by a vC(Ω)Ck,p(Ω)v\in C^{\infty}(\Omega)\cap C^{k,p}(\Omega). This will then establish the claim. For each kNk\in\mathbb{N}, let Ωk=B(0,k){xΩ:dist(x,Ω)>1k}\Omega_{k}=B(0,k)\cap \{x\in\Omega:\dist(x,\partial\Omega)>\frac{1}{k}\}. Clearly Ω=kΩk\Omega=\bigcup_{k}\Omega_{k} and ΩkΩk+1Ω\Omega_{k}\Subset \Omega_{k+1}\Subset\Omega for each kk. Now set Ω0:=\Omega_{0}:=\emptyset and

Ui:=Ωi+1\Ωi1U_{i}:=\Omega_{i+1}\backslash\overline{\Omega}_{i-1}

for each iNi\in\mathbb{N}. We now have that i=1kUi=Ωk+1\bigcup_{i=1}^{k}U_{i}=\Omega_{k+1} so that {Ui}i=1\{U_{i}\}_{i=1}^{\infty} forms an open cover of Ω\Omega. Moreover, clearly UiΩj=U_{i}\cap \Omega_{j}=\emptyset for all ji1j\leq i-1, which implies that {Ui}i=1\{U_{i}\}_{i=1}^{\infty} is a locally finite cover of Ω\Omega. Let {ψi}i=1\{\psi_{i}\}_{i=1}^{\infty} be a partition of unity subordinate to this cover (with the same index).

By Remark 2.4, Jδψiuδ0ψiuJ_{\delta}\psi_{i}\cdot u\xrightarrow{\delta\searrow 0}\psi_{i}\cdot u in Wk,p(Ω)W^{k,p}(\Omega) so that for each ε>0\eps>0 and iNi\in \mathbb{N}, we may choose a δi<1(i+1)(i+2)<dist(Ωi+1,Ω)\delta_{i}<\frac{1}{(i+1)(i+2)}<\dist(\Omega_{i+1},\partial\Omega) such that

Jδi(ψiu)ψiuk,p<ε2i+1.||J_{\delta_{i}}(\psi_{i}\cdot u) - \psi_{i}\cdot u||_{k,p}<\frac{\eps}{2^{i+1}}.

This choice of δi\delta_{i} ensures that supp Jδ(ψiu)Ωi+2\Ωi2\supp J_{\delta}(\psi_{i}\cdot u)\subset \Omega_{i+2}\backslash\Omega_{i-2} so that for each jNj\in\mathbb{N} we have the following equalities when restricting to Ωj\Omega_{j}:

u=i=1ψiu=i=1j+2ψiuv:=i=1Jδi(ψiu)=i=1j+2Jδi(ψiu)\begin{aligned} u&=\sum_{i=1}^{\infty}\psi_{i}\cdot u=\sum_{i=1}^{j+2}\psi_{i}\cdot u\\ v&:=\sum_{i=1}^{\infty}J_{\delta_{i}}(\psi_{i}\cdot u)=\sum_{i=1}^{j+2}J_{\delta_{i}}(\psi_{i}\cdot u) \end{aligned}

Hence, for each jNj\in\mathbb{N}, the triangle inequality implies that

uvk,p,Ωji=1j+2Jδi(ψiu)ψiuk,p<ε2.||u- v||_{k,p,\Omega_{j}} \leq \sum_{i=1}^{j+2}||J_{\delta_{i}}(\psi_{i}\cdot u)-\psi_{i}\cdot u||_{k,p}< \frac{\eps}{2}.

By the monotone convergence theorem, we may take the limit of both sides as jj\rightarrow\infty to obtain uvk,pε2<ε||u-v||_{k,p}\leq \frac{\eps}{2}<\eps.

Remark 2.9. The above claim is false for p=p=\infty for much the same reason LL^{\infty} functions cannot be approximated by continuous functions.

Exercise 2.10. uW1,1(]a,b[)u\in W^{1,1}(]a,b[) iff uu' exists almost everywhere, uL1(]a,b[)u'\in L^{1}(]a,b[) and the fundamental theorem of calculus holds, i.e. for any x,y]a,b[x,y\in\left]a,b\right[ with x<yx<y,

u(x)u(y)=yxu.u(x)-u(y)=\int_{y}^{x}u'.

The case Ω=Rn\Omega=\mathbb{R}^{n} is particularly special.

Theorem 2.11. If p<p<\infty, then Wk,p(Rn)=H0k,p(Rn)W^{k,p}(\mathbb{R}^{n})= H_{0}^{k,p}(\mathbb{R}^{n}).

Proof. Let vC(Rn)Wk,p(Rn)v\in C^{\infty}(\mathbb{R}^{n})\cap W^{k,p}(\mathbb{R}^{n}) be such that uvk,p<ε2||u-v||_{k,p}< \frac{\eps}{2}. Let φ:Rn[0,1]\varphi:\mathbb{R}^{n}\rightarrow [0,1] be a smooth function such that φ1\varphi\equiv 1 on B(0,1)B(0,1) and φ0\varphi \equiv 0 outside B(0,2)B(0,2). For each δ>0\delta>0, define the function

vδ(x):=v(x)φ(δx).v_{\delta}(x):=v(x)\cdot \varphi(\delta\cdot x).

Clearly vδC0(Rn)v_{\delta}\in C_{0}^{\infty}(\mathbb{R}^{n}). Set Vδ:={xRn:x>1δ}V_{\delta}:=\{x\in\mathbb{R}^{n}:|x|>\frac{1}{\delta}\}. The triangle inequality implies that

vvδk,p=vvδk,p,Vδvk,p,Vδ+vδk,p,Vδ.||v-v_{\delta}||_{k,p}=||v-v_{\delta}||_{k,p,V_{\delta}}\leq ||v||_{k,p,V_{\delta}} + ||v_{\delta}||_{k,p,V_{\delta}}.

Now, by the product rule,

Dαvδ=βα(αβ)Dβvδαβ(Dαβφ)(δx)CβαDβv,|\D^{\alpha}v_{\delta}|=\left|\sum_{\beta\leq \alpha}\left(\begin{matrix}\alpha\\ \beta \end{matrix} \right)\D^{\beta}v\cdot \delta^{|\alpha-\beta|}(\D^{\alpha-\beta}\varphi)(\delta x) \right| \leq C\cdot \sum_{\beta\leq\alpha}|\D^{\beta}v|,

where CC is a constant depending only on α\alpha, where we assume that δ1\delta\leq 1. From the above and another application of the triangle inequality, we are left with the inequality vvδk,pC0vk,p,Vδ||v-v_{\delta}||_{k,p}\leq C_{0}\cdot ||v||_{k,p,V_{\delta}}. However, if wLp(Rn)w\in L^{p}(\mathbb{R}^{n}), then by the monotone convergence theorem,

Rnwp=limδ0Rn\Vδwplimδ0Vδwp.\int_{\mathbb{R}^{n}}|w|^{p}=\lim_{\delta\searrow 0}\int_{\mathbb{R}^{n}\backslash V_{\delta}}|w|^{p}\Leftrightarrow \lim_{\delta\searrow 0}\int_{V_{\delta}}|w|^{p}.

Therefore, we may choose δ\delta sufficiently small so that vk,p,Vδ<ε2C0||v||_{k,p,V_{\delta}}<\frac{\eps}{2C_{0}}. This then implies that

uvδk,puvk,p+vvδk,p<ε.||u-v_{\delta}||_{k,p}\leq ||u-v||_{k,p} + ||v-v_{\delta}||_{k,p} < \eps.