Keywords

2010 Mathematics Subject Classification

1 Sobolev Spaces, Inequalities, Dirichlet, and Neumann Problems for the Laplacian

1.1 Sobolev Spaces

Let us introduce the following Sobolev spaces: for any 1 < p < 

$$\displaystyle \begin{aligned}W^{m,p} (\Omega) = \left\{u \in \mathscr{D}^\prime (\Omega);\; \forall \, |\alpha| \leq m,\, D^\alpha u \in L^p (\Omega)\right\} \end{aligned}$$

and

$$\displaystyle \begin{aligned}W^{s,p} (\Omega) = \left\{u \in W^{m,p}(\Omega); \; \int_\Omega \int_\Omega \frac{\left| D^\alpha u(x) - D^\alpha u(y)\right|{}^p}{|x-y|{}^{N + \sigma p}} < \infty,\, \forall \, |\alpha| = m \right\}, \end{aligned}$$

where \(m \in \mathbb {N}, s = m + \sigma , 0 < \sigma < 1\) and Ω is an open set of \(\mathbb {R}^N.\) Equipped with the graph norm, they are Banach spaces.

When \(\Omega = \mathbb {R}^N\), using the Fourier transform, we define for any real number s the space

$$\displaystyle \begin{aligned}H^s (\mathbb{R}^N) = \left\{u \in \mathscr{S}^\prime (\mathbb{R}^N); \int_{\mathbb{R}^N} (1 + |\xi|{}^2)^s \, |\hat{u}(\xi)|{}^2 \, d\xi < \infty \right\}, \end{aligned}$$

which is an Hilbert space for the norm:

$$\displaystyle \begin{aligned}\Vert u\Vert_{H^s(\mathbb{R}^N)} = \left(\int_{\mathbb{R}^N} (1+|\xi|{}^2)^{s}\vert\hat{u}\vert^2 dx\right)^{1/2}. \end{aligned}$$

By Plancherel’s theorem we prove that \(W^{s,2}(\mathbb {R}^N) = H^s(\mathbb {R}^N)\) for all s ≥ 0 and this identity is algebraical and topological. So, in the case p = 2, we denote more simply the space W s, 2( Ω) by H s( Ω).

Definition 1.1

For s > 0 and 1 ≤ p < , we denote

$$\displaystyle \begin{aligned}W_0^{s,p}(\Omega) = \overline{\mathscr{D}(\Omega)}^{\left\| \cdot \right\|{}_{W^{s,p}(\Omega)}}, \end{aligned}$$

and its topological dual space

$$\displaystyle \begin{aligned}W^{-s,p^\prime}(\Omega) = \left[W^{s,p}_0 (\Omega)\right]^\prime, \end{aligned}$$

where p is the conjugate of p: 1∕p + 1∕p  = 1. For p = 2, we will write \(H_0^{s}(\Omega )\) and H s( Ω), respectively.

Proposition 1.2

Suppose \(T \in \mathscr {D}^\prime (\Omega )\) . Then \(T \in W^{-m,p^\prime } (\Omega )\) , with \(m \in \mathbb {N}^*\) , if and only if

$$\displaystyle \begin{aligned}T = \displaystyle \sum_{|\alpha| \leq m} D^\alpha f_\alpha,\quad \mathrm{with}\quad f_\alpha \in L^{p^\prime}(\Omega). \end{aligned}$$

1.2 First Properties

It will be assumed from now on that Ω is a bounded open subset of \(\mathbb {R}^N\) with a Lipschitz boundary.

Let us consider the following space

$$\displaystyle \begin{aligned}\mathscr{D}(\overline{\Omega}) = \left\{v_{|\Omega}; \; v \in \mathscr{D}(\mathbb{R}^N) \right\}. \end{aligned}$$

Theorem 1.3

  1. (i)

    The space \(\mathscr {D}(\overline {\Omega })\) is dense in W s, p( Ω) for any s > 0 (even if Ω is unbounded).

  2. (ii)

    The space \(\mathscr {D}(\mathbb {R}^N)\) is dense in \(W^{s,p}(\mathbb {R}^N)\) for any \(s \in \mathbb {R}\).

As consequence, we have the following property: for any s > 0

$$\displaystyle \begin{aligned}W^{s,p}_0 (\mathbb{R}^N) = W^{s,p}(\mathbb{R}^N)\quad \mathrm{and}\quad W^{-s,p^\prime} (\mathbb{R}^N ) = \left[W^{s,p}(\mathbb{R}^N)\right]^\prime. \end{aligned}$$

But in general, for any s > 0, we have \(W^{s,p}_0 (\Omega ) \subsetneq W^{s,p} (\Omega ).\)

Definition 1.4

For s > 0, we set

$$\displaystyle \begin{aligned}\widetilde{W}^{s,p} (\Omega) = \left\{u \in W^{s,p}(\Omega); \; \widetilde{u} \in W^{s,p}(\mathbb{R}^N) \right\}, \end{aligned}$$

where \(\widetilde {u}\) is the extension by 0 of u outside of Ω.

The space \(\widetilde {W}^{s,p}(\Omega )\) is a Banach space for the norm

$$\displaystyle \begin{aligned}\left\|u\right\|{}_{\widetilde{W}^{s,p}(\Omega)} = \left\| \widetilde{u}\right\|{}_{W^{s,p}(\mathbb{R}^N)}. \end{aligned}$$

It is easy to verify that for any nonnegative integer m

$$\displaystyle \begin{aligned} W^{m,p}_0 (\Omega) \hookrightarrow \widetilde{W}^{m,p}(\Omega) \end{aligned} $$
(1)

and for any \(u \in W^{m,p}_0 (\Omega )\) we have

$$\displaystyle \begin{aligned} \left\|u\right\|{}_{\widetilde{W}^{m,p}(\Omega)} = \left\| u\right\|{}_{W^{m,p}(\Omega)}. \end{aligned} $$
(2)

When s = m + σ with 0 < σ < 1, we can show that

$$\displaystyle \begin{aligned} \left\|u\right\|{}_{\widetilde{W}^{s,p}(\Omega)} \simeq \left\|u\right\|{}_{W^{s,p}(\Omega)} + \displaystyle \sum_{|\alpha| = m} \left\| \frac{D^\alpha u}{\varrho^\sigma} \right\|{}_{L^p(\Omega)}, \end{aligned} $$
(3)

where ϱ(x) = d(x, Γ) and Γ =  Ω.

Theorem 1.5

The space \(\mathscr {D}(\Omega )\) is dense in \(\widetilde {W}^{s,p}(\Omega )\) for all s > 0 (even if Ω is unbounded).

From (1), (2) and the definition of \(W^{m,p}_0(\Omega )\), we deduce the following: for any \( m \in \mathbb {N}^*,\)

$$\displaystyle \begin{aligned} \widetilde{W}^{m,p}(\Omega) = W^{m,p}_0(\Omega). \end{aligned} $$
(4)

Theorem 1.6

For any 0 < s ≤ 1∕p, the space \(\mathscr {D}(\Omega )\) is dense in W s, p( Ω), which means that

$$\displaystyle \begin{aligned} W^{s,p}_0(\Omega) = W^{s,p}(\Omega). \end{aligned} $$
(5)

Theorem 1.7

Let 0 < s ≤ 1 and \(u \in W^{s,p}_0(\Omega ).\) Then

$$\displaystyle \begin{aligned}\frac{u}{\varrho^s} \in L^p (\Omega) \Longleftrightarrow s \neq 1/p \end{aligned}$$

and in this case

$$\displaystyle \begin{aligned}\left\|\frac{u}{\varrho^s}\right\|{}_{L^p(\Omega)} \leq C \, \vert u \vert_{W^{s,p}(\Omega)}, \end{aligned}$$

where the notation |⋅| denotes the semi-norm of W s, p( Ω).

The case s = 1 is known as Hardy’s inequality: for all \(u \in W^{1,p}_0 (\Omega ),\)

$$\displaystyle \begin{aligned}\left\|\frac{u}{\varrho}\right\|{}_{L^p(\Omega)} \leq C \, \left\|\nabla u\right\|{}_{L^p(\Omega)}. \end{aligned}$$

Using again a Hardy’s inequality, we prove the following result:

Theorem 1.8

Let s > 0 and \(u \in W_0^{s,p}(\Omega ).\) Then for any |α|≤ s, we have

$$\displaystyle \begin{aligned} \frac{D^\alpha u}{\varrho^{s - |\alpha|}} \in L^p (\Omega) \Longleftrightarrow s - 1/p \notin \mathbb{N}. \end{aligned} $$
(6)

From (3) and (6), we deduce the following identity:

$$\displaystyle \begin{aligned} \widetilde{W}^{s,p}(\Omega) = W^{s,p}_0(\Omega) \end{aligned} $$
(7)

which holds for any s > 0 satisfying \(s - 1/p \notin \mathbb {N}.\)

Proposition 1.9

  1. (i)

    For any 1 ≤ j  N and for any \(s \in \mathbb {R}\) , the operator

    $$\displaystyle \begin{aligned} \frac{\partial }{\partial x_j} : W^{s,p}(\mathbb{R}^N) \longrightarrow W^{s-1,p}(\mathbb{R}^N) \end{aligned} $$
    (8)

    is continuous.

  2. (ii)

    However, if we replace \(\mathbb {R}^N\) by Ω, Property (8) takes place unless s = 1∕p.

Sketch of the Proof of Point (ii)

1. Case :

s = m + σ, with \(m\in \mathbb {N}^*\) and 0 ≤ σ < 1. Let u ∈ W s, p( Ω). By definition, we know that

$$\displaystyle \begin{aligned}u \in W^{m,p}(\Omega) \;\; \; \mbox{and} \; \; \; \int_\Omega \int_\Omega \frac{\left| D^\alpha u(x) - D^\alpha u(y)\right|{}^p}{|x-y|{}^{N + \sigma p}} <\infty, \quad \forall \, |\alpha| = m. \end{aligned}$$

So for any 1 ≤ j ≤ N

$$\displaystyle \begin{aligned}\frac{\partial u}{\partial x_j} \in W^{m-1,p}(\Omega) \quad \mbox{and} \quad \int_\Omega \int_\Omega \frac{\left| D^\alpha \frac{\partial u}{\partial x_j}(x) - D^\alpha \frac{\partial u}{\partial x_j}(y)\right|{}^p}{|x-y|{}^{N + \sigma p}} <\infty, \end{aligned}$$

for all |α| = m − 1. Consequently \(\frac {\partial u}{\partial x_j} \in W^{s-1,p}(\Omega ).\)

2. Case s ≤ 0.:

Let u ∈ W s, p( Ω). Since − s + 1 ≥ 1, for any \(\varphi \in \mathscr {D}(\Omega )\), we get:

$$\displaystyle \begin{aligned}\begin{array}{ll} \left|\langle \frac{\partial u}{\partial x_j},\varphi\rangle_{ \mathscr{D}^\prime (\Omega) \times \mathscr{D}(\Omega)} \right| &= \left|- \langle u,\frac{\partial \varphi}{\partial x_j}\rangle_{\mathscr{D}^\prime (\Omega) \times \mathscr{D}(\Omega)} \right| \\ &\leq \left\| u\right\|{}_{W^{s,p}(\Omega)} \left\| \frac{\partial \varphi}{\partial x_j}\right\|{}_{W_0^{-s,p^\prime}(\Omega)}\\ &\leq \left\| u\right\|{}_{W^{s,p}(\Omega)} \left\| \varphi\right\|{}_{W_0^{-s +1,p^\prime}(\Omega)}. \end{array} \end{aligned}$$

We conclude by using the density of \(\mathscr {D}(\Omega )\) in \(W^{-s +1,p^\prime }_0(\Omega )\).

3. Case 0 < s < 1.:

Let u ∈ W s, p( Ω). Recall that Ω being Lipschitz open set, there exists an extension operator

$$\displaystyle \begin{aligned}\forall t \geq 0, \quad P : W^{t,p}(\Omega) \longrightarrow W^{t,p}(\mathbb{R}^N) \end{aligned}$$

which is linear, continuous, and satisfying

$$\displaystyle \begin{aligned}Pv_{|\Omega} = v, \quad \mbox{for any }\; v \in W^{t,p}(\Omega). \end{aligned}$$

As \(Pu \in W^{s,p}(\mathbb {R}^N)\), we get \(\frac {\partial Pu}{\partial x_j} \in W^{s-1,p} (\mathbb {R}^N).\) But

$$\displaystyle \begin{aligned}\left(\frac{\partial Pu}{\partial x_j} \right)_{|\Omega} = \frac{\partial u}{\partial x_j}, \end{aligned}$$

where \(\frac {\partial u}{\partial x_j}\) is the restriction to Ω of the distribution \(T = \frac {\partial Pu}{\partial x_j} \in W^{s-1,p}(\mathbb {R}^N)\). More precisely, we have:

$$\displaystyle \begin{aligned}\forall \, \varphi \in \mathscr{D}(\Omega), \quad \langle \frac{\partial u}{\partial x_j},\varphi \rangle_{\mathscr{D}^\prime(\Omega) \times \mathscr{D}(\Omega)} = \langle T, \widetilde{\varphi}\rangle_{\mathscr{D}^\prime(\mathbb{R}^N) \times \mathscr{D}(\mathbb{R}^N)}. \end{aligned}$$

That implies

$$\displaystyle \begin{aligned}\left| \langle \frac{\partial u}{\partial x_j},\varphi \rangle \right| \leq \left\|T\right\|{}_{W^{s-1,p}(\mathbb{R}^N)} \left\|\widetilde{\varphi}\right\|{}_{W^{1-s,p^\prime} (\mathbb{R}^N)} = \left\|T\right\|{}_{W^{s-1,p}(\mathbb{R}^N)} \left\|\varphi\right\|{}_{\widetilde{W}^{1-s,p^\prime} (\Omega)}. \end{aligned}$$

We have shown that \(\frac {\partial u}{\partial x_j} \in \left [\widetilde {W}^{1-s,p^\prime }(\Omega )\right ]^\prime .\) But

$$\displaystyle \begin{aligned}\left[\widetilde{W}^{1-s,p^\prime}(\Omega)\right]^\prime = \left[W_0^{1-s,p^\prime}(\Omega)\right]^\prime \Longleftrightarrow 1-s \neq 1/p^\prime, \end{aligned}$$

i.e., s ≠ 1∕p. □

Remark 1

The above proof shows that

$$\displaystyle \begin{aligned}u \in W^{1/p,p}(\Omega) \Longrightarrow \frac{\partial u}{\partial x_j} \in \left[\widetilde{W}^{1/p^\prime,p^\prime}\right]^\prime. \end{aligned}$$

In particular,

$$\displaystyle \begin{aligned}u \in H^{1/2}(\Omega) \Longrightarrow \frac{\partial u}{\partial x_j}\in \left[\widetilde{H}^{1/2}(\Omega)\right]^\prime, \end{aligned}$$

where we remark also that

$$\displaystyle \begin{aligned}\widetilde{H}^{1/2}(\Omega) \hookrightarrow H^{1/2}(\Omega) = H^{1/2}_0 (\Omega). \end{aligned}$$

This embedding being dense, we get by duality

$$\displaystyle \begin{aligned}H^{-1/2} (\Omega) = \left[H^{1/2}_0 (\Omega)\right]^\prime \hookrightarrow \left[\widetilde{H}^{1/2}(\Omega)\right]^\prime . \end{aligned}$$

Corollary 1.10

Let s > 0. The following characterization holds:

$$\displaystyle \begin{aligned}u \in \widetilde{W}^{s,p}(\Omega) \Longleftrightarrow u \in W^{s,p}_0(\Omega) \quad \mathit{\mbox{and}} \quad \mathit{\mbox{for any}} \; |\alpha| = m, \; \; \frac{D^\alpha u}{\varrho^\sigma} \in L^p (\Omega), \end{aligned}$$

where \(s = m + \sigma , m \in \mathbb {N}\) and 0 ≤ σ < 1.

1.3 Traces

Firstly, recall the following inclusions:

$$\displaystyle \begin{aligned}W^{s,p}(\mathbb{R}^N) \hookrightarrow \mathscr{C}^0 (\mathbb{R}^N) \quad \mbox{if} \quad s >\frac{N}{p}. \end{aligned}$$

So that if \(u \in W^{s,p}(\mathbb {R}^N)\) with \(s >\frac {N}{p}\), the restriction of u to the hyperplane x N = 0 is well defined. But the continuity with respect to all variables is not necessary. It is enough to have the continuity with respect to the variable x N. This is possible as soon as s > 1∕p.

Actually, we have the following result:

Theorem 1.11

  1. (i)

    Suppose that s − 1∕p = k + σ, with \( k \in \mathbb {N}\) and 0 < σ < 1 (which implies, in particular, that \(s- 1/p \notin \mathbb {N}\) ). Then the mapping

    where

    $$\displaystyle \begin{aligned}\gamma_0 u (x) = u(x^\prime,0), x^\prime =(x_1,\ldots, x_{N-1}), \quad and\quad \gamma_j u (x^\prime) = \frac{\partial^j u}{\partial x_N^j} (x^\prime,0), \end{aligned}$$

    defined for \(u \in \mathscr {D}(\mathbb {R}^N)\) , has a unique extension

    $$\displaystyle \begin{aligned}W^{s,p}(\mathbb{R}^n) \longrightarrow \displaystyle \prod_{j=0}^k W^{s-j-1/p,p}(\mathbb{R}^{N-1}) \end{aligned}$$

    which is continuous and where k is the integer part of s > 0.

  2. (ii)

    Moreover this operator has a right continuous inverse R:

    $$\displaystyle \begin{aligned}\left\{ \begin{array}{ll} \forall \, \boldsymbol{g} = (g_0,\ldots, g_k) \in \displaystyle \prod_{j=0}^k W^{s-j-1/p,p}(\mathbb{R}^{N-1}), \quad \gamma R \boldsymbol{g} = \boldsymbol{g} \\ \left\| R \boldsymbol{g} \right\|{}_{W^{s,p}(\mathbb{R}^N)} \leq C_N \, \displaystyle \sum_{j=0}^k \left\|g_j\right\|{}_{W^{s-j -1/p,p}(\mathbb{R}^{N-1})}. \end{array} \right. \end{aligned}$$

Remark 2

For p = 2, the above result can be proved using the Fourier transform.

This result can be extended to the case where Ω is a bounded open subset of \(\mathbb {R}^{N}\), with a \(\mathscr {C}^{k, 1} \) boundary (see the definition below).

Definition 1.12

Let Ω be an open subset of \(\mathbb {R}^N\). We say that Ω is Lipschitz (respectively of class \(\mathscr {C}^{k,1}, k \in \mathbb {N}^\star \)) if for every x ∈ Γ, there exists a neighborhood V of x in \(\mathbb {R}^N\) and orthonormal coordinates \(\left \{y_1,\ldots ,y_N\right \}\) satisfying:

  1. (i)

    V is an hypercube

    $$\displaystyle \begin{aligned}V = \left\{(y_1,\ldots,y_N) \in \mathbb{R}^N; \; |y_j| <a_j, \; 1 \leq j \leq N \right\}, \end{aligned}$$
  2. (ii)

    there exists a function φ defined in

    $$\displaystyle \begin{aligned}V^\prime = \left\{ y^\prime\in \mathbb{R}^{N-1}; \; |y_j| < a_j, 1 \leq j \leq N-1 \right\}, \end{aligned}$$

    such that φ and φ −1 are Lipschitz (respectively, \(\mathscr {C}^{k,1}\)) and satisfying (Fig. 1)

    $$\displaystyle \begin{aligned}\forall \, y^\prime \in V^\prime , \quad \left|\varphi(y^\prime)\right| \leq \frac{1}{2} a_N \end{aligned}$$
    figure 1

    Fig. 1

    $$\displaystyle \begin{aligned}\Omega \cap V = \left\{(y^\prime,y_N) \in V;\; y_N < \varphi (y^\prime) \right\} \end{aligned}$$
    $$\displaystyle \begin{aligned}\Gamma \cap V = \left\{(y^\prime,y_n) \in V; \; y_N = \varphi(y^\prime) \right\}. \end{aligned}$$

Let

$$\displaystyle \begin{aligned}\begin{array}{ll} \Phi : & V^\prime \longrightarrow \Gamma \cap V\\ & y^\prime \longmapsto (y^\prime,\varphi(y^\prime)). \end{array} \end{aligned}$$

Definition 1.13

Suppose that Ω is an open subset of \(\mathbb {R}^N\) of class \(\mathscr {C}^{k,1}\), with \(k \in \mathbb {N}\) and let 0 < s ≤ k + 1. We introduce the following space

$$\displaystyle \begin{aligned}W^{s,p}(\Gamma) = \left\{u \in L^p(\Gamma);\; u\circ \Phi \in W^{s,p}(V^\prime \cap \Phi^{-1}(\Gamma \cap V)) \right\} \end{aligned}$$

for any (V, φ) verifying the previous definition.

Let (V j, φ j), 1 ≤ j ≤ J, be any atlas of Γ for which each pair (V j, φ j) satisfies the above definition. One possible Banach norm for W s, p( Γ) is given by:

$$\displaystyle \begin{aligned}\left\|u\right\|{}_{W^{s,p}(\Gamma)} = \displaystyle \sum_{j=1}^J \left\|u\circ \Phi_j\right\|{}_{W^{s,p}(V^\prime_j \cap \Phi_j^{-1} (\Gamma \cap V_j))} \end{aligned}$$

which is equivalent when 0 < s < 1 to the norm

$$\displaystyle \begin{aligned}\left(\left\|u\right\|{}^p_{L^p(\Gamma)} + \int_\Gamma \int_\Gamma \frac{\left|u(x) - u(y)\right|{}^p}{|x-y|{}^{N-1 + sp}} d\sigma_x d\sigma_y \right)^{1/p}. \end{aligned}$$

We are now in position to extend Theorem 1.11 to the case where \(\mathbb {R}^{N-1}\) is replaced by an N − 1-dimensional manifold of \(\mathbb {R}^N\), but which is sufficiently regular. This simply uses changes of variables.

If locally Γ is represented by the pair (V, φ) with φ and φ −1 Lipschitz, then a unit outward normal vector can be defined as follows:

$$\displaystyle \begin{aligned}\mbox{for}\; y^\prime \in V^\prime,\quad \boldsymbol{\nu}(y^\prime,\varphi(y^\prime)) = \frac{(-\nabla^\prime \varphi(y^\prime),1)}{\sqrt{1+ |\nabla^\prime \varphi (y^\prime)|{}^2}}. \end{aligned}$$

One can then extend this vector in all V by setting

$$\displaystyle \begin{aligned}\boldsymbol{\nu} (y^\prime,y_N) =\boldsymbol{\nu}(y^\prime,\varphi(y^\prime)), y \in V. \end{aligned}$$

As \(\Gamma \subset \displaystyle \cup _{j=1}^J V_j\), we know that there exist functions \(\mu _0, \mu _1,\ldots , \mu _J \in \mathscr {C}^\infty (\mathbb {R}^N)\) such that

  1. (i)

    for all j = 0, …, J, \(0 \leq \mu _j \leq 1 \quad \mbox{and} \quad \displaystyle \sum _{j=1}^J \mu _j = 1\)

  2. (ii)

    supp μ j is compact and supp μ j ⊂ V j for any j ≥ 1 and supp μ 0 ⊂ Ω.

This partition of unity then allows to extend ν in a neighborhood of \(\overline {\Omega }\) as follows: \(\boldsymbol {\nu } = \displaystyle \sum _{j=0}^J (\mu _j\boldsymbol {\nu }).\) It is then easy to verify that \(\boldsymbol {\nu } \in L^\infty (\overline {\Omega })\) if Γ is Lipschitz and \(\boldsymbol {\nu } \in \mathscr {C}^{k-1,1}(\overline {\Omega })\) if Γ is \(\mathscr {C}^{k,1}\).

We are now ready to establish the following result:

Theorem 1.14 (Traces)

Let Ω be an open subset of \(\mathbb {R}^N\) of class \(\mathscr {C}^{k,1}\) , with \(k \in \mathbb {N}\) . Let s > 0 satisfying s  k + 1 and s − 1∕p = ℓ + σ with 0 < σ < 1 and \(\ell \in \mathbb {N}.\) Then the mapping

defined for \(\mathscr {C}^{k,1}\) has a unique continuous extension as an operator from W s, p( Ω) into \( \displaystyle \prod _{j=0}^\ell W^{s-j - 1/p,p} (\Gamma )\) where

$$\displaystyle \begin{aligned}\gamma_1 u = \frac{\partial u}{\partial\boldsymbol{\nu}} = \nabla u \cdot \boldsymbol{\nu}, \quad \gamma_j u = \frac{\partial^j u}{\partial\boldsymbol{\nu}^j}. \end{aligned}$$

Moreover this operator has a right continuous inverse R (not depending of p).

Case Ω Lipschitz. Suppose 1∕p < s ≤ 1. We have the following properties:

  1. (i)

    If u ∈ W s, p( Ω), then u | Γ∈ W s−1∕p, p( Γ).

  2. (ii)

    If g ∈ W s−1∕p, p( Γ), then there exists u ∈ W s, p( Ω) such that u = g on Γ and satisfying the estimate

    $$\displaystyle \begin{aligned}\left\|u\right\|{}_{W^{s,p}(\Omega)} \leq C \, \left\|g \right\|{}_{W^{s - 1/p,p}(\Gamma)}. \end{aligned}$$

Case Ω of class \(\mathscr {C}^{1,1}\) .

  1. (i)

    Let u ∈ W s, p( Ω). If 1∕p < s ≤ 2, then u | Γ∈ W 1−1∕p( Γ). Moreover, for any g ∈ W s−1∕p, p( Γ), there exists u ∈ W s, p( Ω) such that u = g on Γ, with

    $$\displaystyle \begin{aligned}\left\|u\right\|{}_{W^{s,p}(\Omega)} \leq C \, \left\|g\right\|{}_{W^{s-1/p,p}(\Gamma)}. \end{aligned}$$
  2. (ii)

    Let u ∈ W s, p( Ω). If 1 + 1∕p < s ≤ 2, then \(\frac {\partial u}{\partial \boldsymbol {\nu }} \in W^{s-1-1/p,p}(\Gamma ).\) Moreover, for any g 0 ∈ W s−1∕p, p( Γ) and g 1 ∈ W s−1−1∕p, p( Γ), there exists u ∈ W s, p( Ω) such that

    $$\displaystyle \begin{aligned}u = g_0 \quad \mbox{and} \quad \frac{\partial u}{\partial\boldsymbol{\nu}} = g_1 \quad \mbox{on} \; \Gamma \end{aligned}$$

    with

    $$\displaystyle \begin{aligned}\left\|u\right\|{}_{W^{s,p}(\Omega)} \leq C \, \left( \left\|g_0\right\|{}_{W^{s-1/p,p}(\Gamma)} + \left\|g_1\right\|{}_{W^{s-1-1/p,p}(\Gamma)}\right). \end{aligned}$$

Theorem 1.15

Suppose that Ω is an open subset of \(\mathbb {R}^N\) of class \(\mathscr {C}^{k,1}\) , with \(k \in \mathbb {N}\) . Let s > 0 such that \(s- 1/p \notin \mathbb {N}\) and s − 1∕p = ℓ + σ, where 0 < σ < 1 and ℓ ≥ 0 is an integer. Then we have the following characterization for s  k + 1:

$$\displaystyle \begin{aligned}W^{s,p}_0(\Omega) = \left\{u \in W^{s,p}(\Omega); \gamma_0 u = \gamma_1 u = \ldots = \gamma_\ell u = 0 \right\}. \end{aligned}$$

1.4 Interpolation

We will consider here only the case of spaces H s( Ω), with Ω bounded open Lipschitz of \(\mathbb {R}^N\).

Recall that for every s > 0 there exists a continuous linear operator:

$$\displaystyle \begin{aligned}P : H^s (\Omega) \longrightarrow H^s (\mathbb{R}^N) \end{aligned}$$

satisfying

$$\displaystyle \begin{aligned}\, \forall \, u \in H^s(\Omega), \quad Pu_{|\Omega} = u.\end{aligned} $$

Theorem 1.16

[Interpolation Inequality] Let s 1, s 2, s 3 with 0 ≤ s 1 < s 2 < s 3. Then

$$\displaystyle \begin{aligned}\forall \, \varepsilon > 0,\quad \left\|u\right\|{}_{W^{s_2,p}(\Omega)}\leq \varepsilon \left\|u\right\|{}_{W^{s_3,p}(\Omega)} + K\varepsilon^{- \frac{s_2-s_1}{s_3 - s_2}} \left\| u\right\|{}_{W^{s_1,p}(\Omega)},\end{aligned} $$

where K = K( Ω, s 1, s 2, s 3, p).

The above inequality is a consequence of the compactness of the embedding of \(W^{s_3,p}(\Omega )\) into \(W^{s_2,p}(\Omega )\).

Recall now that we have different ways to define the Sobolev space H m( Ω), for \(m \in \mathbb {N}\):

$$\displaystyle \begin{aligned} \begin{array}{ll} u \in H^m(\Omega) \iff \forall \, |\alpha| \leq m, \; D^\alpha u \in L^2(\Omega),\\ u \in H^m (\Omega)\iff u = U_{|\Omega} \; \mbox{with} \; U \in H^m (\mathbb{R}^N),\\ u \in H^m (\mathbb{R}^N) \iff u\in \mathscr{S}^\prime (\mathbb{R}^N) \quad \mathrm{and}\quad (1 + |\xi|{}^2)^{m/2} \hat{u} \in L^2(\mathbb{R}^N). \end{array} \end{aligned} $$
(9)

In the case of fractional Sobolev spaces H s( Ω), with \(s = m + \sigma , m\in \mathbb {N}, 0 < \sigma < 1\), we have:

$$\displaystyle \begin{aligned} \begin{array}{ll} u \in H^s(\Omega) \iff u \in H^m(\Omega) \quad \mathrm{and}\quad \forall \, |\alpha| = m,\; \int_\Omega \int_\Omega \frac{\left|D^\alpha u(x) - D^\alpha u(y)\right|}{|x-y|{}^{N + 2 \sigma}} < \infty\\ u \in H^s (\Omega) \iff u = U_{|\Omega} \;\quad \mathrm{with} \quad U \in H^s (\mathbb{R}^N), \\ u \in H^s (\mathbb{R}^N) \iff u\in \mathscr{S}^\prime (\mathbb{R}^N) \quad \mathrm{and}\quad (1 + |\xi|{}^2)^{s/2} \hat{u} \in L^2(\mathbb{R}^N). \end{array} \end{aligned} $$
(10)

We can also get this space by interpolation:

$$\displaystyle \begin{aligned}H^s(\Omega) = \left[H^m(\Omega), L^2(\Omega)\right]_{\mu}, \, 0 <\mu < 1 \quad (1-\mu) m = s\end{aligned} $$

and more generally we have for any 0 < μ < 1

$$\displaystyle \begin{aligned}\left[H^{s_1}(\Omega), H^{s_2}(\Omega)\right]_{\mu} = H^{(1-\mu)s_1 + \mu s_2} (\Omega).\end{aligned} $$

Concerning the interpolation of spaces \(H^m_0(\Omega )\), we have:

$$\displaystyle \begin{aligned}\left[H^{s_1}_0(\Omega), H^{s_2}_0(\Omega)\right]_{\mu} = H_0^{(1-\mu)s_1 + \mu s_2} (\Omega)\quad \mbox{ if }\quad (1-\mu) s_1 + \mu s_2\notin \frac{1}{2} + \mathbb{N}\end{aligned} $$

and

$$\displaystyle \begin{aligned}\left[H_0^{s_1}(\Omega), H_0^{s_2}(\Omega)\right]_{\mu} = \widetilde{H}^{(1-\mu)s_1 + \mu s_2} (\Omega)\quad \mathrm{otherwise}, \end{aligned}$$

with equivalent norms.

1.5 Transposition

Let V and H be two Hilbert spaces on \(\mathbb {R}\) and \(A \in \mathscr {L} (V,H)\). For every fixed g ∈ H , we consider the following mapping

$$\displaystyle \begin{aligned}\begin{array}{ll} V & \longrightarrow \mathbb{R} \\ x & \longmapsto \langle g,Ax\rangle_{H^\prime \times H} \end{array}\end{aligned} $$

which defines a linear and continuous form on V that we denote by t Ag:

$$\displaystyle \begin{aligned}\langle ^t A g, x\rangle_{V^\prime \times V} = \langle g,Ax\rangle_{H^\prime \times H}. \end{aligned}$$

Remark 3

If A : V →H is an isomorphism, then we can define the transpose of A −1 and we easily verify that

$$\displaystyle \begin{aligned}^t{A^{-1}} = \left({}^t A\right)^{-1} \quad \mbox{and} \quad ^t A : H^\prime \longrightarrow V^\prime \; \mbox{is an isomorphism}.\end{aligned} $$

1.6 Inequalities

They are fundamental tools in the study of partial differential equations:

  1. (i)

    Poincaré’s Inequality. Let Ω be an open space bounded in at least one direction. Then there exists a constant C ≥ 0, depending on the diameter of Ω such that

    $$\displaystyle \begin{aligned}\forall \, u \in W^{1,p}_0 (\Omega),\quad \left\|u\right\|{}_{L^p(\Omega)} \leq C \, \left\|\nabla u\right\|{}_{L^p(\Omega)}.\end{aligned} $$
  2. (ii)

    Poincaré-Wirtinger’s Inequality. Let Ω be a Lipschitz bounded domain of \(\mathbb {R}^N\). Then there exists a constant C( Ω) ≥ 0 such that

    $$\displaystyle \begin{aligned}\forall \, u \in W^{1,p}(\Omega),\quad \displaystyle \inf_{K \in \mathbb{R}} \left\|u + K\right\|{}_{L^p(\Omega)} \leq C(\Omega) \, \left\|\nabla u \right\|{}_{L^p(\Omega)}. \end{aligned}$$
  3. (iii)

    Hardy’s Inequality. Let Ω be a Lipschitz bounded open subset of \(\mathbb {R}^N\). Then there exists a constant C( Ω) ≥ 0 such that

    $$\displaystyle \begin{aligned}\forall \, u \in W^{1,p}_0 (\Omega), \quad \left\|\frac{u}{\varrho}\right\|{}_{L^p(\Omega)} \leq C(\Omega) \, \left\|\nabla u \right\|{}_{L^p(\Omega)}. \end{aligned}$$
  4. (iv)

    Calderòn–Zygmund’s Inequality.

    $$\displaystyle \begin{aligned}\forall \, u \in \mathscr{D}(\Omega), \quad \left\|\frac{\partial^2 u}{\partial x_i \partial x_j}\right\|{}_{L^p(\Omega)} \leq C(\Omega) \, \left\|\Delta u \right\|{}_{L^p(\Omega)}. \end{aligned}$$

1.7 Weak Solutions

Consider the following problems:

$$\displaystyle \begin{aligned}(P_D) \quad - \Delta u = f \quad \mbox{in} \; \Omega \quad \quad \mathrm{and}\quad u = g \quad \mbox{on} \; \Gamma \end{aligned}$$

and

$$\displaystyle \begin{aligned}(P_N) \quad - \Delta u = f \quad \mbox{in} \; \Omega \quad \quad \mathrm{and}\quad \frac{\partial u}{\partial\boldsymbol{\nu}} = h \quad \mbox{on} \; \Gamma , \end{aligned}$$

where Ω is a Lipschitz bounded domain of \(\mathbb {R}^N\), f, g, and h are given.

Theorem 1.17

Given any f  H −1( Ω) and any g  H 1∕2( Γ), there exists a unique solution u  H 1( Ω) to Problem (P D). Moreover

$$\displaystyle \begin{aligned}\left\|u\right\|{}_{H^1(\Omega)} \leq C(\Omega) \, \left(\left\|f\right\|{}_{H^{-1}(\Omega)} + \left\|g\right\|{}_{H^{1/2}(\Gamma)}\right). \end{aligned}$$

Proof

Using Theorem 1.14, there exists u g ∈ H 1( Ω) such that

$$\displaystyle \begin{aligned}u_g = g \quad \mbox{on} \; \Gamma \quad \mathrm{with} \quad \left\|u_g\right\|{}_{H^1(\Omega)} \leq C(\Omega) \, \left\|g \right\|{}_{H^{1/2}(\Gamma)}. \end{aligned}$$

Setting

$$\displaystyle \begin{aligned}f_g = - \Delta u_g = - \mathrm{div}\, \nabla u_g \in H^{-1}(\Omega), \end{aligned}$$

the problem becomes: Find \(v\in H^1_0(\Omega )\) solution of

$$\displaystyle \begin{aligned}(P^0_D) \quad - \Delta v = f - f_g \quad \mbox{in} \; \Omega \quad \quad \mathrm{and}\quad v = 0 \quad \mbox{on} \; \Gamma . \end{aligned}$$

This last problem is equivalent to the following variational formulation:

$$\displaystyle \begin{aligned}(FV)_D \, \left\{ \begin{array}{ll} \mbox{Find} \; v \in H^1_0(\Omega) \; \mbox{such that} \\ \forall \, \varphi \in H^1_0(\Omega), \quad \displaystyle \int_\Omega \nabla v \cdot \nabla \varphi dx = \langle f- f_g, \varphi \rangle_{{H^{-1}(\Omega} \times H^1_0(\Omega)}. \end{array} \right. \end{aligned}$$

Applying Lax–Milgram Lemma or Riesz Theorem, we prove the existence of a unique solution \(v \in H^1_0(\Omega ) \) satisfying (FV )D.

Note that the bilinear form

$$\displaystyle \begin{aligned}a(v,\varphi) = \int_\Omega \nabla v \cdot \nabla \varphi dx \end{aligned}$$

is continuous on \(H^1_0(\Omega ) \times H^1_0(\Omega )\) and coercive on \(H^1_0(\Omega )\) thanks to Poincaré’s inequality. In addition, this form allows to define a scalar product on Hilbert’s space \(H^1_0(\Omega )\). □

Remark 4

  1. (i)

    If Ω is of class \(\mathscr {C}^{1}\), f ∈ W −1, p( Ω) and g ∈ W 1−1∕p, p( Γ) with 1 < p < , then there exists a unique solution u ∈ W 1, p( Ω) to (P D).

  2. (ii)

    When Ω is only Lipschitz, this regularity result holds for p ∈ ]2 − ε , 2 + ε[ where ε and ε  > 0 are depending on Ω and 2 − ε and 2 + ε are conjugate.

Concerning the Neumann problem, the approach is a bit more complicated. Indeed, if we are looking for a solution u ∈ H 1( Ω) only, the boundary condition on the normal derivative does not make sense, since the functions of L 2( Ω) do not have any trace at the boundary. Here, in fact, if one set v = ∇u we have

$$\displaystyle \begin{aligned}\frac{\partial u}{\partial\boldsymbol{\nu}} = \boldsymbol{v} \cdot \boldsymbol{\nu} \; \mbox{on} \; \Gamma. \end{aligned}$$

Definition 1.18

$$\displaystyle \begin{aligned}H(\mathrm{div};\, \Omega) = \left\{ \boldsymbol{v} \in L^2(\Omega); \; \mathrm{div}\, \boldsymbol{v} \in L^2(\Omega) \right\}. \end{aligned}$$

It is a Hilbert space for the scalar product

$$\displaystyle \begin{aligned}(( \boldsymbol{v},\boldsymbol{w}))_{H(\mathrm{div};\, \Omega)} = \int_\Omega \boldsymbol{v} \cdot \boldsymbol{w} dx + \int_\Omega (\mathrm{div}\, \boldsymbol{v}) (\mathrm{div}\, \boldsymbol{w}) dx. \end{aligned}$$

Proposition 1.19

  1. (i)

    The space \(\mathscr {D}(\overline {\Omega })\) is dense in H(div;  Ω).

  2. (ii)

    The linear mapping

    $$\displaystyle \begin{aligned}\boldsymbol{v} \longmapsto \boldsymbol{v}\cdot \boldsymbol{\nu}, \end{aligned}$$

    defined on \(\mathscr {D}(\overline {\Omega })^N,\) can be uniquely extended into a linear mapping of H(div; Ω) in \(H^{-1/2}(\Gamma ):= \left [H^{1/2}(\Gamma )\right ]^\prime \).

  3. (iii)

    In addition, we have the following Green’s formula (or Stokes’ formula):

    $$\displaystyle \begin{aligned}\forall \varphi \in H^1(\Omega),\; \forall \boldsymbol{v}\in H(\mathrm{div}; \, \Omega),\quad \int_\Omega \boldsymbol{v} \cdot \nabla \varphi \, dx + \int_\Omega \varphi \, \mathrm{div}\, \boldsymbol{v} \, dx = \langle \boldsymbol{v}\cdot\boldsymbol{\nu},\varphi \rangle_{\Gamma} \end{aligned}$$

    where 〈⋅, ⋅〉Γ denotes the duality brackets H −1∕2( Γ) × H 1∕2( Γ).

Corollary 1.20

Let u  H 1( Ω) be such that Δu  L 2( Ω). Then \(\frac {\partial u}{\partial \boldsymbol {\nu }} \in H^{-1/2} (\Gamma ).\) Moreover for any φ  H 1( Ω), we have the following Green formula:

$$\displaystyle \begin{aligned}\int_\Omega \varphi \Delta u \, dx + \int_\Omega \nabla u \cdot \nabla \varphi \, dx = \langle \frac{\partial u}{\partial\boldsymbol{\nu}}, \varphi \rangle_\Gamma. \end{aligned}$$

Proof

It suffices to apply Proposition 1.19 by setting v = ∇u. □

As a Consequence we can show that for any f ∈ L 2( Ω) and for any g ∈ H −1∕2( Γ), the problems

$$\displaystyle \begin{aligned}(P_N) \, \left\{ \begin{array}{ll} \mbox{Find} \; u \in H^1(\Omega)\; \mbox{such that} \\ -\Delta u = f \quad \mbox{in} \; \Omega, \\ \frac{\partial u}{\partial\boldsymbol{\nu}} = g \quad \mbox{on} \; \Gamma \end{array} \right. \end{aligned}$$

and

$$\displaystyle \begin{aligned}(Q_N) \, \left\{ \begin{array}{ll} \mbox{Find} \; u \in H^1(\Omega) \; \mbox{such that} \\ \forall \, \varphi \in H^1(\Omega),\quad \displaystyle \int_\Omega \nabla u \cdot \nabla \varphi \, dx = \int_\Omega f \varphi \, dx + \langle g, \varphi \rangle_{\Gamma} \end{array} \right. \end{aligned}$$

are equivalent, so that any solution of one is a solution of the other.

Remark 5

  1. (i)

    The open Ω being bounded, the constant functions belong to H 1( Ω). So that if u is a solution of (Q N), taking φ = 1, the data f and g must satisfy the (necessary) compatibility condition:

    $$\displaystyle \begin{aligned}\int_\Omega f \, dx + \langle g,1\rangle_\Gamma = 0. \end{aligned}$$
  2. (ii)

    The implication (P N)⇒(Q N) results from Corollary 1.20. The reverse implication also uses Green’s formula and the surjectivity of the trace operator of H 1( Ω) into H 1∕2( Γ).

Theorem 1.21

Let Ω be a bounded, connected, and Lipschitzian open of \(\mathbb {R}^N,\) with N ≥ 2. Let f  L 2( Ω), g  H −1∕2( Γ) satisfying the compatibility condition

$$\displaystyle \begin{aligned}\int_\Omega f \, dx + \langle g,1\rangle_\Gamma = 0. \end{aligned}$$

Then Problem (P N) has a solution H 1( Ω), unique to an additive constant, verifying the estimate:

$$\displaystyle \begin{aligned}\left\|\nabla u\right\|{}_{L^2(\Omega)} \leq C(\Omega) \, \left(\left\|f \right\|{}_{L^2(\Omega)} + \left\|g\right\|{}_{H^{-1/2}(\Gamma)} \right). \end{aligned}$$

Proof

According to Poincaré-Wirtinger’s inequality, we have

$$\displaystyle \begin{aligned}\displaystyle \inf_{K \in \mathbb{R}} \left\|u + K \right\|{}_{H^1(\Omega)} \leq C(\Omega) \, \left\|\nabla u \right\|{}_{L^2(\Omega)}. \end{aligned}$$

So that the bilinear form

$$\displaystyle \begin{aligned}a(u,\varphi) = \int_\Omega \nabla u \cdot\nabla \varphi \, dx \end{aligned}$$

is coercive on the quotient space \(V = H^1(\Omega ) /_{ \displaystyle {\mathbb {R}}}\). It is then sufficient to apply Lax–Milgram on the Hilbert space V . □

Remark 6

  1. (i)

    We could have chosen as space V the space \(H^1(\Omega ) \cap L^2_0(\Omega )\) where

    $$\displaystyle \begin{aligned}L^2_0(\Omega) = \left\{v \in L^2(\Omega); \int_\Omega v\, dx = 0 \right\}, \end{aligned}$$

    which is a Hilbert space and then use the inequality:

    $$\displaystyle \begin{aligned}\forall \, v \in H^1(\Omega) \cap L^2_0(\Omega), \quad \left\|v\right\|{}_{H^1(\Omega)} \leq C \, \left\| \nabla v\right\|{}_{L^2(\Omega)}. \end{aligned}$$
  2. (ii)

    We could have taken f in a space larger than L 2( Ω). More precisely if \(f \in L^{(2^*)^\prime }(\Omega )\), where (2) is the conjugate of 2 defined by

    $$\displaystyle \begin{aligned}\frac{1}{2^*} = \left\{ \begin{array}{ll} \frac{1}{2} - \frac{1}{N} \quad \mbox{if} \; N \geq 3 \\ \varepsilon > 0 \quad \mbox{arbitrary if} \;N =2, \end{array} \right. \end{aligned}$$

    i.e., \((2^*)^\prime = \frac {2N}{N + 2}\) if N ≥ 3 and (2) > 1 if N = 2.

  3. (iii)

    In L p-theory, we have existence results in W 1, p( Ω) when Ω is \(\mathscr {C}^{1}\) and 1 < p <  or when Ω is \(\mathscr {C}^{0,1}\) and 2 − ε  < p < 2 + ε.

In the same spirit, we can consider the case of Fourier-Robin boundary condition:

$$\displaystyle \begin{aligned}(P_{FR}) \left\{ \begin{array}{ll} \mbox{Find} \; u \in H^1(\Omega) \\ - \Delta u = f \quad \mbox{in}\; \Omega, \\ \frac{\partial u}{\partial\boldsymbol{\nu}} + \alpha u = g \quad \mbox{on } \; \Gamma, \end{array} \right. \end{aligned}$$

where α is a positive function defined on Γ, which can be formulated in an equivalent way by:

$$\displaystyle \begin{aligned}(Q_{FR}) \left\{ \begin{array}{ll} \mbox{Find} \; u \in H^1(\Omega) \; \mbox{such that} \\ \forall \, \varphi \in H^1(\Omega), \quad \displaystyle \int_\Omega \nabla u \cdot \nabla \varphi \, dx + \int_\Gamma \alpha u \varphi \, dx = \int_\Omega f \varphi \, dx + \langle g,\varphi \rangle_\Gamma. \end{array} \right. \end{aligned}$$

1.8 Strong Solutions

Theorem 1.22

Let Ω be a bounded open of class \(\mathscr {C}^{1,1}\) of \(\mathbb {R}^N\) . Let f  L 2( Ω) and g  H ∕2( Γ). Then the solution u given by Theorem 1.17 belongs to H 2( Ω) and verifies the estimate:

$$\displaystyle \begin{aligned}\left\|u\right\|{}_{H^2(\Omega)} \leq C(\Omega) \, \left( \left\|f\right\|{}_{L^2( \Omega)} + \left\|g\right\|{}_{H^{3/2}(\Gamma)} \right). \end{aligned}$$

Proof

Firstly, we note that

$$\displaystyle \begin{aligned}L^2(\Omega) \hookrightarrow H^{-1}(\Omega) \quad \mbox{and} \quad H^{3/2}(\Gamma) \hookrightarrow H^{1/2}(\Gamma) \end{aligned}$$

so that the problem (P D) has a unique solution u ∈ H 1( Ω).

We shift the data g ∈ H 3∕2( Γ) by u g ∈ H 2( Ω) and we set again u = v + u g, so that v ∈ H 1( Ω) vérifies:

$$\displaystyle \begin{aligned}\left\{ \begin{array}{ll} - \Delta v = f + \Delta u_g \in L^2(\Omega), \\ v = 0 \quad \mbox{on} \; \Gamma. \end{array} \right. \end{aligned}$$

So, we need to show that v ∈ H 2( Ω). One of the methods to establish this regularity consists in using the technique of the differential quotients.

The complete proof being long and tedious, we will admit it. □

Remark 7

We can also establish the existence of solutions in W 2, p( Ω) when the data f and g verify:

$$\displaystyle \begin{aligned}f \in L^p(\Omega)\quad \mathrm{and} \quad g \in W^{2 - 1/p,p}(\Gamma) \end{aligned}$$

and the domain Ω is of class \(\mathscr {C}^{1,1}\).

1.9 Very Weak Solutions

We assume here that Ω is a bounded open of class \(\mathscr {C}^{1,1}\) and we are interested in the homogeneous problem

$$\displaystyle \begin{aligned}(P^H_{D}) \left\{ \begin{array}{ll} \mbox{Find} \; u \in L^2(\Omega) \\ - \Delta u = 0 \quad \mbox{in}\; \Omega, \\ u = g \quad \mbox{on } \; \Gamma, \end{array} \right. \end{aligned}$$

where g ∈ H −1∕2( Γ).

Remark 8

As the function u belongs “only” to L 2( Ω), the boundary condition u = g on Γ has a priori no sense. But we will see that in fact, we can make sense of the trace of a harmonic function in L 2( Ω) and (we can in fact weaken this last hypothesis).

Lemma 1.23

  1. (i)

    The space \(\mathscr {D}(\overline {\Omega })\) is dense in the space

    $$\displaystyle \begin{aligned}E(\Omega;\Delta) = \left\{v \in L^2(\Omega);\; \Delta v \in L^2(\Omega) \right\}. \end{aligned}$$
  2. (ii)

    The mapping vv | Γ defined on \(\mathscr {D}(\overline {\Omega })\) can be uniquely extended into a continuous linear mapping of E( Ω; Δ) into H −1∕2( Γ).

  3. (iii)

    In addition, we have the following Green’s formula:

    $$\displaystyle \begin{aligned}\left\{ \begin{array}{ll} \forall \, v \in E(\Omega;\Delta),\quad \forall \, \varphi \in H^2(\Omega) \cap H^1_0(\Omega) \\ \displaystyle \int_\Omega v \Delta \varphi \, dx - \int_\Omega \varphi \Delta v\, dx = \langle v, \frac{\partial \varphi}{\partial\boldsymbol{\nu}} \rangle_{H^{-1/2}(\Gamma) \times H^{1/2}(\Gamma)}. \end{array} \right. \end{aligned}$$

Proof

  1. (i)

    The idea is to use the Hahn–Banach theorem. So let \(\ell \in \left [E(\Omega ;\Delta )\right ]^\prime \) vanishing on \(\mathscr {D}(\overline {\Omega })\) and show that it cancels on E( Ω; Δ).

    We know that there exist (f, g) ∈ L 2( Ω) × L 2( Ω) such that

    $$\displaystyle \begin{aligned}\forall \, v \in E(\Omega;\Delta), \quad \langle \ell,v\rangle = \int_\Omega f v \, dx + \int_\Omega g \Delta v \, dx. \end{aligned}$$

    Let \(\widetilde {f}\) and \(\widetilde {g}\) the extensions by 0 outside of Ω of f and g, respectively. Then, for any \(v\in \mathscr {D}(\mathbb {R}^N)\)

    $$\displaystyle \begin{aligned}\langle \ell,\, v_{\vert\Omega} \rangle = \int_\Omega fv \, dx + \int_\Omega g \Delta v \, dx = \int_{\mathbb{R}^N} \widetilde{f} v \, dx + \int_{\mathbb{R}^N} \widetilde{g} \Delta v \, dx, \end{aligned}$$

    i.e.,

    $$\displaystyle \begin{aligned}\Delta \widetilde{g} = - \widetilde{f} \; \mbox{in} \; \mathbb{R}^N. \end{aligned}$$

    As \(\widetilde {g} \in L^2(\mathbb {R}^N)\) and \(\Delta \widetilde {g} \in L^2(\mathbb {R}^N),\) then \(\widetilde {g} \in H^2(\mathbb {R}^N)\). Therefore, g ∈ H 2( Ω). The extension \(\widetilde {g}\), by 0 outside of Ω, belongs to \(H^2(\mathbb {R}^N)\). We know then that \(g \in H^2_0(\Omega )\). By definition, there exists a sequence (g k)k of functions of \(\mathscr {D}(\Omega )\) such that g kg in H 2( Ω).

    Finally, let v ∈ E( Ω; Δ). So,

    $$\displaystyle \begin{aligned}\langle \ell,v \rangle = \displaystyle\lim_{k \Longrightarrow \infty} \left[ \int_\Omega - v \Delta v_k \, dx + \int_\Omega g_k \Delta v \, dx \right] = \displaystyle \lim_{k \Longrightarrow \infty} 0 = 0. \end{aligned}$$
  2. (ii)

    Let \(v \in \mathscr {D}(\overline {\Omega })\) fixed and \(\varphi \in H^2(\Omega ) \cap H^1_0 (\Omega )\). Then

    $$\displaystyle \begin{aligned}\int_\Omega v \Delta \varphi \, dx - \int_\Omega \varphi \Delta v \, dx = \int_\Gamma v \frac{\partial \varphi }{\partial\boldsymbol{\nu}}. \end{aligned}$$

    Now let μ ∈ H 1∕2( Γ). According to the trace theorem and since Ω is of class \(\mathscr {C}^{1,1}\), there exists φ ∈ H 2( Ω) verifying

    $$\displaystyle \begin{aligned}\left\{ \begin{array}{ll} \varphi = 0 \quad \mbox{and} \quad \frac{\partial \varphi}{\partial\boldsymbol{\nu}} = \mu \quad \mbox{on} \; \Gamma, \\ \left\|\varphi\right\|{}_{H^2(\Omega)} \leq C \, \left\|\mu\right\|{}_{H^{1/2}(\Gamma)}. \end{array} \right. \end{aligned}$$

    Thus, using the Cauchy–Schwarz inequality

    $$\displaystyle \begin{aligned}\left|\langle v,\mu\rangle_{H^{-1/2}(\Gamma) \times H^{1/2}(\Gamma)} \right| = \left| \int_\Gamma v \mu \right| = \left|\int_\Gamma v \frac{\partial \varphi}{\partial\boldsymbol{\nu}}\right| \end{aligned}$$
    $$\displaystyle \begin{aligned}\leq C(\Omega) \, \left( \left\|v\right\|{}^2_{L^2(\Omega)} + \left\|\Delta v \right\|{}_{L^2(\Omega)}^2 \right)^{1/2} \, \left\|\varphi\right\|{}_{H^2(\Omega)} \end{aligned}$$
    $$\displaystyle \begin{aligned}\leq C(\Omega) \left\|v\right\|{}_{E(\Omega;\Delta)} \left\|\mu\right\|{}_{H^{1/2}(\Gamma)}. \end{aligned}$$

    This shows that the linear mapping

    $$\displaystyle \begin{aligned}\begin{array}{ll} \mathscr{D}(\overline{\Omega}) & \longrightarrow H^{-1/2}(\Gamma) \\ \; \; \; \; v & \longmapsto v_{|\Gamma} \end{array} \end{aligned}$$

    is continuous when \(\mathscr {D}(\overline {\Omega })\) is equipped with the norm of E( Ω; Δ). We finish the proof by using the density of \(\mathscr {D}(\overline {\Omega })\) in E( Ω; Δ).

  3. (iii)

    Immediate.

Theorem 1.24

Let Ω be a bounded open of class \(\mathscr {C}^{1,1}\) of \(\mathbb {R}^N\) and let g  H −1∕2( Γ). Then, the problem \((P_D^0)\) has a unique solution u  L 2( Ω) verifying the estimate

$$\displaystyle \begin{aligned}\left\|u\right\|{}_{L^2(\Omega)} \leq C(\Omega) \, \left\|g\right\|{}_{H^{-1/2}(\Gamma)}. \end{aligned}$$

Proof

From Green’s formula above, it is easy to see that u ∈ L 2( Ω) is a solution of the problem \((P_D^0)\) if and only if

$$\displaystyle \begin{aligned} \forall \, \varphi \in H^2(\Omega) \cap H^1_0(\Omega), \quad \displaystyle \int_\Omega u \Delta \varphi \, dx = \langle g,\frac{\partial \varphi}{\partial\boldsymbol{\nu}} \rangle_\Gamma. \end{aligned} $$
(11)

Indeed, let u ∈ L 2( Ω) be a solution of \((P_D^0)\). Green’s formula implies that (11) takes place.

Conversely, let u ∈ L 2( Ω) be a solution of (11). Then, for all \(\varphi \in \mathscr {D}(\Omega )\), we have

$$\displaystyle \begin{aligned}0 = \int_\Omega u \Delta \varphi \, dx = \langle \Delta u, \varphi\rangle_{\mathscr{D}^\prime (\Omega) \times \mathscr{D}(\Omega)}, \end{aligned}$$

i.e.,

$$\displaystyle \begin{aligned} \Delta u = 0 \quad \mbox{in} \; \Omega. \end{aligned} $$
(12)

Let now \(\varphi \in H^2(\Omega ) \cap H^1_0(\Omega ).\) From (12) and Green’s formula above, we deduce successively that:

$$\displaystyle \begin{aligned}0 =\displaystyle \int_\Omega \varphi \Delta u \, dx = \int_\Omega u \Delta \varphi\, dx - \langle u,\frac{\partial \varphi}{\partial\boldsymbol{\nu}} \rangle_\Gamma \end{aligned}$$

then

$$\displaystyle \begin{aligned}\langle u,\frac{\partial \varphi}{\partial\boldsymbol{\nu}}\rangle_\Gamma = \langle g,\frac{\partial \varphi}{\partial\boldsymbol{\nu}}\rangle_\Gamma. \end{aligned}$$

From the surjectivity of the trace mapping \(v \mapsto (v_{\vert \Gamma }, \frac {\partial v}{\partial \boldsymbol {\nu }})\) from H 2( Ω) into H 3∕2( Γ) × H 1∕2( Γ) we know that

$$\displaystyle \begin{aligned}\forall \mu \in H^{1/2}(\Gamma), \quad \langle u,\mu \rangle_\Gamma = \langle g,\mu \rangle_\Gamma, \end{aligned}$$

i.e., u = g in H −1∕2( Γ). □

Remark 9

A similar result can be established for the Neumann problem \((P_N^0)\) with boundary data h in H −3∕2( Γ) and satisfying the compatibility condition 〈h, 1〉Γ = 0.

1.10 Solutions in H s( Ω), with 0 < s < 2

We have established in the previous paragraphs the existence of solutions in H 1( Ω), H 2( Ω), and L 2( Ω) under generally optimal assumptions (except for the Neumann problem).

We will now consider the case of solutions in H s( Ω) with 0 < s < 2 and s ≠ 1. The main ingredient is to use interpolation (complex here).

Theorem 1.25

Let Ω be a bounded open of class \(\mathscr {C}^{1,1}\).

  1. (i)

    Suppose that \(\frac {1}{2} <s < 2.\) Then the operators

    $$\displaystyle \begin{aligned} \begin{array}{rl} & \Delta : H^s(\Omega) \cap H^1_0(\Omega) \longrightarrow H^{s-2}(\Omega) = \left[H_0^{2-s}(\Omega)\right]^\prime\quad \mathrm{if}\ \; 1 < s < 2 \; \mathrm{and}\; s \neq \frac{3}{2},\\ &\Delta : H^{3/2}_0(\Omega) \longrightarrow \left[H_{00}^{1/2}(\Omega)\right]^\prime, \\ & \Delta : H^{2-s}_0(\Omega) \longrightarrow H^{-s}(\Omega) = \left[H_0^{s}(\Omega)\right]^\prime \quad \mathrm{if}\ \; 1 < s < \frac{3}{2}, \end{array} \end{aligned} $$
    (13)

    are isomorphisms.

  2. (ii)

    For any g  H s( Γ), with \(-\frac {1}{2} < s < \frac {3}{2}\) , Problem \((P_D^H)\) has a unique solution \(u \in H^{s + \frac {1}{2}} (\Omega )\).

Remark 10

What happens if Ω is only Lipschitz? For what values of s can we have u ∈ H s( Ω)?

2 The Stokes Problem with Various Boundary Conditions

We are interested here in the study of the Stokes problem:

$$\displaystyle \begin{aligned}(S) \, \left\{ \begin{array}{ll} \mbox{Find} \; (\boldsymbol{u}, \pi) \quad \mbox{satisfying} \\ - \Delta \boldsymbol{u} + \nabla \pi = \boldsymbol{f}\quad \mbox{in} \; \Omega, \\ \mathrm{div} \, \boldsymbol{u} = 0 \quad \mbox{in} \; \Omega, \end{array} \right. \end{aligned}$$

with one of the following boundary conditions on Γ:

  1. (i)

    u = 0 (Dirichlet boundary condition)

  2. (ii)

    u ⋅ν = 0 and curl u ×ν = 0 (Navier type boundary condition)

  3. (iii)

    \(\boldsymbol {u} \cdot \boldsymbol {\nu } = 0 \; \mbox{and} \; (\mathbb {D}\boldsymbol {u})\boldsymbol {\nu } + \alpha \boldsymbol {u}_\tau = \mathbf {0} \; (\mbox{Navier boundary condition})\)

  4. (iv)

    u ×ν = 0 and π = π 0 (pressure boundary condition).

Here u denotes the velocity field, π the pressure field, Ω a connected bounded open set we assume at least Lipschitz.

Recall that

$$\displaystyle \begin{aligned}\mathrm{div}\ \, \boldsymbol{u} = \nabla \cdot \boldsymbol{u}, \quad \mathbf{curl}\, \boldsymbol{u} = \nabla \times \boldsymbol{u}\quad \mathrm{and}\quad \mathbb{D}\boldsymbol{u} = \frac{1}{2} \left( \nabla \boldsymbol{u} + (\nabla \boldsymbol{u})^T\right). \end{aligned}$$

The notation u τ denotes the tangential component of u: u τ = u − (u ⋅ν)ν. Finally f and α are given on Ω and Γ, respectively.

Remark 11

  1. (i)

    We limit ourselves here, with the exception of pressure, to the case of homogeneous boundary conditions.

  2. (ii)

    If the boundary of Ω is flat (like a cube, for example, or half space), the above boundary conditions are more easily written. When \(\Omega = \mathbb {R}_+^3\), the Navier type boundary condition is equivalent to:

    $$\displaystyle \begin{aligned}u_3 = 0 \quad \mbox{and} \quad \frac{\partial u_1}{\partial x_3} = \frac{\partial u_2}{\partial x_3} = 0 \end{aligned}$$

    and that of Navier at:

    $$\displaystyle \begin{aligned}u_3 = 0 \quad \mbox{and} \quad \frac{\partial u_1}{\partial x_3} - \alpha u_1 = \frac{\partial u_2}{\partial x_3} - \alpha u_2 = 0. \end{aligned}$$

2.1 The Problem (S) with Dirichlet Boundary Condition

As for the Laplace equation with the Dirichlet boundary condition, we will assume

$$\displaystyle \begin{aligned}\boldsymbol{f}\in H^{-1}(\Omega)^3 \end{aligned}$$

and so look for \(\boldsymbol {u} \in H^1_0(\Omega )^3\) verifying (S). Here we have in addition the constraint

$$\displaystyle \begin{aligned}\mathrm{div}\, \boldsymbol{u} = 0 \quad \mbox{in} \; \Omega \end{aligned}$$

and the Lagrange multiplier π. First of all, as π must verify

$$\displaystyle \begin{aligned}\nabla \pi = \boldsymbol{f}+ \Delta \boldsymbol{u} \in H^{-1}(\Omega)^3 \end{aligned}$$

it is, therefore, reasonable to look for π in L 2( Ω). Moreover, it is easy to verify that such π satisfies:

$$\displaystyle \begin{aligned}\forall \, \boldsymbol{v} \in H^1_0(\Omega)^3,\quad \langle \nabla \pi,\boldsymbol{v} \rangle_{H^{-1}(\Omega) \times H^1_0(\Omega)} = - \int_\Omega \pi \, \mathrm{div}\, \boldsymbol{v} \, dx. \end{aligned}$$

The space

$$\displaystyle \begin{aligned}V = \left\{\boldsymbol{v} \in H^1_0(\Omega)^3; \; \mathrm{div}\, \boldsymbol{v} = 0 \; \mbox{in} \; \Omega \right\} \end{aligned}$$

being a subspace of \(H^1_0(\Omega )^3\) is, therefore, a Hilbert space. Moreover

$$\displaystyle \begin{aligned}\forall \, \boldsymbol{v} \in V, \quad \langle \nabla \pi,\boldsymbol{v} \rangle_{H^{-1}(\Omega) \times H^1_0(\Omega)} = 0. \end{aligned}$$

We are now able to propose a variational formulation of Problem (S):

$$\displaystyle \begin{aligned}(P_D^0) \, \left\{ \begin{array}{ll} \mbox{Find} \; \boldsymbol{u} \in V \; \mbox{such that} \\ \displaystyle \forall \, \boldsymbol{v} \in V,\quad \int_\Omega \nabla \boldsymbol{u} : \nabla \boldsymbol{v} \, dx = \langle \boldsymbol{f},\boldsymbol{v} \rangle_{H^{-1}(\Omega) \times H^1_0(\Omega)}, \end{array} \right. \end{aligned}$$

where we note that the pressure π has “disappeared.”

Lemma 2.1

The problem

$$\displaystyle \begin{aligned}(S_D^0) \, \left\{ \begin{array}{ll} \mathit{\mbox{Find}} \; (\boldsymbol{u}, \pi) \in H^1_0(\Omega)^3 \times L^2(\Omega)\\ - \Delta \boldsymbol{u} + \nabla \pi = \boldsymbol{f}\quad \mathit{\mbox{in}} \; \Omega, \\ \mathrm{div} \, \boldsymbol{u} = 0 \quad \mathit{\mbox{in}} \; \Omega \end{array} \right. \end{aligned}$$

is equivalent to the problem \((P_D^0)\).

Proof

The implication \((S_D^0) \Longrightarrow (P_D^0)\) is immediate. Conversely, let u be a solution of \((P_D^0)\). Then, in particular,

$$\displaystyle \begin{aligned}\forall \, \boldsymbol{v} \in \mathscr{D}(\Omega)^3 \quad \mbox{such that} \quad \mathrm{div}\, \boldsymbol{v} = 0 \quad \mbox{in} \; \Omega, \end{aligned}$$

we have

$$\displaystyle \begin{aligned} \langle - \Delta \boldsymbol{u} - \boldsymbol{f}, \boldsymbol{v} \rangle_{\mathscr{D}^\prime(\Omega)^3 \times \mathscr{D}(\Omega)^3} = 0. \end{aligned} $$
(14)

As − Δu −f ∈ H −1( Ω)3 and the space

$$\displaystyle \begin{aligned}\mathcal{V}(\Omega) = \left\{\boldsymbol{v} \in \mathscr{D}(\Omega)^3; \mathrm{div}\, \boldsymbol{v} = 0 \ \mbox{in} \; \Omega \right\} \end{aligned}$$

is dense in the space V , then the relation (14) takes place for all v. Then we know that there exists π ∈ L 2( Ω), unique up to an additive constant, because Ω is connected, such that

$$\displaystyle \begin{aligned}- \Delta \boldsymbol{u} - \boldsymbol{f}= \nabla(-\pi) \quad \mbox{in} \; \Omega \end{aligned}$$

(this result is called “De Rham’s version of the theorem” in H −1( Ω)N). And finally, as u ∈ V , then

$$\displaystyle \begin{aligned}\mathrm{div}\, \boldsymbol{u} = 0 \; \mbox{in} \; \Omega \quad \mbox{and} \quad \boldsymbol{u} = \mathbf{0} \quad \mbox{on} \; \Gamma. \end{aligned}$$

This ends the proof of the lemma. □

Theorem 2.2

For any f ∈ H −1( Ω)3, the Stokes problem \((P_D^0)\) has a unique solution u ∈ V vérifying further

$$\displaystyle \begin{aligned}\left\|\boldsymbol{u}\right\|{}_{H^1(\Omega)^3} \leq C(\Omega) \, \left\|\boldsymbol{f}\, \right\|{}_{H^{-1}(\Omega)^3}. \end{aligned}$$

Proof

Simply apply Lax–Milgram theorem. □

Remark 12

The theory is well known for everything that concerns the regularity of solutions when the data are:

  • solutions in W 1, p( Ω)3 × L p( Ω)

  • solutions in W 2, p( Ω)3 × L p( Ω)

with 1 < p < .

In particular, if f ∈ L 2( Ω)3 and Ω is of class \(\mathscr {C}^{1,1}\), then u ∈ H 2( Ω)3 and π ∈ H 1( Ω).

2.2 The Stokes Problem with Navier Type Boundary Condition

Here we are still interested in Stokes’ problem, but with the following boundary condition:

$$\displaystyle \begin{aligned}\boldsymbol{u} \cdot\boldsymbol{\nu} = 0 \quad \mbox{and} \quad \mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu} = \mathbf{0} \;\mbox{on} \; \Gamma. \end{aligned}$$

In order to take into account this condition at the boundary, it is important to write the Laplacian operator in the form:

$$\displaystyle \begin{aligned}- \Delta = \mathbf{curl}\, \mathbf{curl}\, - \nabla \, \mathrm{div}. \end{aligned}$$

On the other hand, if we study the existence of weak solutions u in H 1( Ω)3, it will be necessary to give a meaning to the condition at the boundary

$$\displaystyle \begin{aligned}\mathbf{curl}\, \boldsymbol{u} \times \boldsymbol{\nu} = \mathbf{0} \quad \mbox{on} \; \Gamma. \end{aligned}$$

Recall the following Green formulas:

  1. (i)

    If v ∈ L 2( Ω)3 and curl v ∈ L 2( Ω)3, then v ×ν ∈ H −1∕2( Γ)3 and

    $$\displaystyle \begin{aligned}\forall \, \boldsymbol{\varphi} \in H^1(\Omega)^3, \quad \int_\Omega \boldsymbol{v} \cdot \mathbf{curl}\, \boldsymbol{\varphi} \, dx - \int_\Omega \boldsymbol{\varphi} \cdot \mathbf{curl}\, \boldsymbol{v} \, dx = \langle \boldsymbol{v} \times\boldsymbol{\nu}, \boldsymbol{\varphi} \rangle_\Gamma, \end{aligned}$$

    where 〈⋅, ⋅〉Γ denotes the duality brackets H −1∕2( Γ) × H 1∕2( Γ).

  2. (ii)

    If v ∈ L 2( Ω)3 and div v ∈ L 2( Ω), then v ⋅ν ∈ H −1∕2( Γ) and

    $$\displaystyle \begin{aligned}\forall \, \varphi \in H^1(\Omega), \quad \int_\Omega \boldsymbol{v} \cdot \nabla \varphi \, dx + \int_\Omega \varphi \, \mathrm{div}\, \boldsymbol{v} \, dx = \langle \boldsymbol{v} \cdot \boldsymbol{\nu}, \varphi \rangle_\Gamma. \end{aligned}$$

Remark 13

If v ∈ L 2( Ω)3 and curl v ∈ L 6∕5( Ω)3 (respectively, div v ∈ L 6∕5( Ω)), then

$$\displaystyle \begin{aligned}\boldsymbol{v} \times\boldsymbol{\nu} \in H^{-1/2}(\Gamma)^3 \; (\mbox{resp.} \; \boldsymbol{v} \cdot \boldsymbol{\nu} \in H^{-1/2}(\Gamma)) \end{aligned}$$

and Green’s formulas above remain valid.

Proposition 2.3

Let v ∈ L 2( Ω)3 such that curl v ∈ L 2( Ω)3 and curl curl v ∈ L 6∕5( Ω)3. Then curl v ×ν ∈ H −1∕2( Γ)3 and we have the following Green formula:

$$\displaystyle \begin{aligned}\forall \, \boldsymbol{\varphi} \in H^1(\Omega)^3, \int_\Omega \mathbf{curl}\, \boldsymbol{v} \cdot\mathbf{curl}\, \boldsymbol{\varphi} - \int_\Omega \boldsymbol{\varphi} \cdot \mathbf{curl}\, \mathbf{curl}\, \boldsymbol{v} = \langle \mathbf{curl}\, \boldsymbol{v} \times\boldsymbol{\nu},\boldsymbol{\varphi} \rangle_\Gamma. \end{aligned}$$

Proof

It suffices to put w = curl v and use the previous reminders. □

We are now able to propose a variational formulation for the Stokes problem (S) with the Navier type homogeneous condition. To do this, we set

$$\displaystyle \begin{aligned}V = \left\{ \boldsymbol{v} \in L^2(\Omega)^3; \mathbf{curl}\, \boldsymbol{v} \in L^2 (\Omega), \mathrm{div}\ \, \boldsymbol{v} = 0 \; \mbox{in} \; \Omega \; \mbox{and} \; \boldsymbol{v} \cdot \boldsymbol{\nu} = 0 \; \mbox{on} \; \Gamma \right\} \end{aligned}$$

equipped with the graph norm:

$$\displaystyle \begin{aligned}\left\| \boldsymbol{v}\right\|{}_V = \left( \left\|\boldsymbol{v}\right\|{}^2_{L^2(\Omega)} + \left\|\mathbf{curl}\, \boldsymbol{v}\right\|{}^2_{L^2(\Omega)^3}\right)^{1/2} \end{aligned}$$

which makes it a Hilbert space.

We suppose f ∈ L 6∕5( Ω)3 and we consider the following variational formulation:

$$\displaystyle \begin{aligned}(P^0_{TN}) \left\{ \begin{array}{ll} \mbox{Find} \; \boldsymbol{u} \in V \; \mbox{such that for any} \; \boldsymbol{v} \in V, \\ \displaystyle \int_\Omega \mathbf{curl}\, \boldsymbol{u} \cdot \mathbf{curl}\, \boldsymbol{v} \, dx = \int_\Omega\boldsymbol{f} \cdot \boldsymbol{v} \, dx. \end{array} \right. \end{aligned}$$

Questions

  1. (i)

    Is the problem \((P_{TN}^0)\) equivalent to the problem \((S^0_{TN})\)?

  2. (ii)

    If so, is the bilinear form

    $$\displaystyle \begin{aligned}\begin{array}{ll} V \times V & \longrightarrow \mathbb{R} \\ (\boldsymbol{u},\boldsymbol{v}) & \longmapsto \displaystyle \int_\Omega \mathbf{curl}\, \boldsymbol{u} \cdot \mathbf{curl}\, \boldsymbol{v} \, dx \end{array} \end{aligned}$$

    coercive?

Remark 14

As with the Neumann problem for the Laplacian, the boundary condition

$$\displaystyle \begin{aligned}\mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu} = \mathbf{0} \quad \mbox{on} \; \Gamma \end{aligned}$$

is “hidden” in the variational formulation.

Answers to the Above Questions

In order to study Problem \((P^0_{TN})\), we have to describe with more precision the geometry of the domain. We first need the following definition.

Definition 2.4

A bounded domain in \(\mathbb {R}^{3}\) is called pseudo-\(\mathscr {C}^{\,0,1}\) (respectively, pseudo-\(\mathscr {C}^{\,1,1}\)) if for any point x on the boundary there exists an integer r(x) equal to 1 or 2 and a strictly positive real number λ 0 such that for all real numbers λ with 0 < λ < λ 0, the intersection of Ω with the ball with center x and radius λ, has r(x) connected components, each one being \(\mathscr {C}^{\,0,1}\) (resp. \(\mathscr {C}^{\,1,1}\)).

Hypothesis

There exist J connected open surfaces Σj, 1 ≤ j ≤ J, called “cuts,” contained in Ω, such that:

  1. (i)

    each surface Σj is an open part of a smooth manifold \( \mathcal {M}_{j}\),

  2. (ii)

    the boundary of Σj is contained in  Ω for 1 ≤ j ≤ J,

  3. (iii)

    the intersection \(\bar {\Sigma _{i}}\cap \bar {\Sigma _{j}}\) is empty for i ≠ j,

  4. (iv)

    the open set

    $$\displaystyle \begin{aligned}\Omega^{\circ}=\Omega\,\backslash \,\bigcup_{j=1}^{J} \Sigma_{j} \end{aligned}$$

    is pseudo-\(\mathscr {C}^{\,0,1}\) (respectively, pseudo-\(\mathscr {C}^{\,1,1}\)) simply connected.

Fig. 2
figure 2

Example for J = 1 and I = 3

Theorem 2.5

Let Ω be a bounded open \(\mathscr {C}^{\,1,1}\) set.

  1. (i)

    Let v ∈ L 2( Ω)3 such that div v ∈ L 2( Ω), curl v ∈ L 2( Ω) and satisfying in addition

    $$\displaystyle \begin{aligned}\boldsymbol{v} \cdot \boldsymbol{\nu} \in H^{1/2}(\Gamma) \, \quad (\mathit{\mbox{respectively,}} \; \boldsymbol{v} \times\boldsymbol{\nu} \in H^{1/2} (\Gamma)^3). \end{aligned}$$

    Then v ∈ H 1( Ω)3 and we have the following estimates:

    $$\displaystyle \begin{aligned} \Vert \boldsymbol{v} \Vert_{H^1(\Omega)} \leq C(\Omega) ( \left\|\boldsymbol{v}\right\|{}_{L^2(\Omega)} + \left\|\mathrm{div}\, \boldsymbol{v} \right\|{}_{L^2(\Omega)} + \left\|\mathbf{curl}\, \boldsymbol{v} \right\|{}_{L^2(\Omega)} + \left\|\boldsymbol{v} \cdot\boldsymbol{\nu} \right\|{}_{H^{1/2}(\Gamma)}){} \end{aligned} $$
    (15)

    and

    $$\displaystyle \begin{aligned} \left\|\boldsymbol{v}\right\|{}_{H^1(\Omega)} \leq C(\Omega) \, \left[ \left\|\boldsymbol{v}\right\|{}_{L^2(\Omega)} + \left\|\mathrm{div}\, \boldsymbol{v} \right\|{}_{L^2(\Omega)} + \left\|\mathbf{curl}\, \boldsymbol{v} \right\|{}_{L^2(\Omega)} + \left\|\boldsymbol{v} \times\boldsymbol{\nu}\right\|{}_{H^{1/2}(\Gamma)}) \right]. \end{aligned} $$
    (16)
  2. (ii)

    Under the above assumptions, if in addition v ⋅ν = 0 on Γ, then we have the following estimate:

    $$\displaystyle \begin{aligned} \left\|\boldsymbol{v}\right\|{}_{H^1(\Omega)} \leq C(\Omega) \, \big( \left\|div \, \boldsymbol{v}\right\|{}_{L^2(\Omega)} + \left\|\mathbf{curl}\, \boldsymbol{v}\right\|{}_{L^2(\Omega)} + \displaystyle\sum_{j=1}^J \left| \int_{\Sigma_j} \boldsymbol{v} \cdot\boldsymbol{\nu}\right|\big) \end{aligned} $$
    (17)

    and if v ×ν = 0 on Γ, then we have the following estimate:

    $$\displaystyle \begin{aligned} \left\|\boldsymbol{v}\right\|{}_{H^1(\Omega)} \leq C(\Omega) \, \big( \left\|div \, \boldsymbol{v}\right\|{}_{L^2(\Omega)} + \left\|\mathbf{curl}\, \boldsymbol{v}\right\|{}_{L^2(\Omega)} + \displaystyle\sum_{i=1}^J \left| \int_{\Gamma_i} \boldsymbol{v} \cdot\boldsymbol{\nu} \right| \big). \end{aligned} $$
    (18)

Remark 15

  1. (i)

    Suppose that

    $$\displaystyle \begin{aligned}\boldsymbol{v} \in L^2(\Omega)^3,\; \mathrm{div} \, \boldsymbol{v} \in L^2(\Omega) \quad \mbox{and} \quad \mathbf{curl}\, \boldsymbol{v} \in L^2(\Omega)^3 \end{aligned}$$

    with

    $$\displaystyle \begin{aligned}\boldsymbol{v} \cdot\boldsymbol{\nu} = 0 \quad \mbox{and} \quad \boldsymbol{v} \times\boldsymbol{\nu} = \mathbf{0} \quad \mbox{on} \; \Gamma. \end{aligned}$$

    Let us then extend v by 0 outside of Ω. It is easy to show that this extension verifies:

    $$\displaystyle \begin{aligned}\widetilde{\boldsymbol{v}} \in L^2(\mathbb{R}^3)^3, \; \mathrm{div}\ \, \widetilde{\boldsymbol{v}} \subset L^2(\mathbb{R}^3) \quad \mbox{and} \quad \mathbf{curl}\, \widetilde{\boldsymbol{v}} \in L^2(\mathbb{R}^3)^3. \end{aligned}$$

    As − Δ = curl curl −∇ div, then \(\Delta \widetilde {\boldsymbol {v}} \in H^{-1}(\mathbb {R}^3)^3\) and

    $$\displaystyle \begin{aligned}\widetilde{\boldsymbol{v}} - \Delta \widetilde{\boldsymbol{v}} \in H^{-1}(\mathbb{R}^3)^3, \end{aligned}$$

    which means that \(\widetilde {\boldsymbol {v}} \in H^1(\mathbb {R}^3)^3\) and, therefore, \(\boldsymbol {v} \in H^1_0(\Omega )^3\).

  2. (ii)

    Now note that if \(\boldsymbol {u} \in \mathscr {D}(\mathbb {R}^3)^3\), then

    $$\displaystyle \begin{aligned} \begin{array}{ll} \displaystyle \int_\Omega \left|\nabla \boldsymbol{u}\right|{}^2 \, dx &= - \displaystyle \int_{\mathbb{R}^3} \boldsymbol{u} \cdot \Delta \boldsymbol{u} \, dx = \displaystyle \int_{\mathbb{R}^3} [\boldsymbol{u} \cdot (\mathbf{curl}\, \mathbf{curl}\, \boldsymbol{u}) - \boldsymbol{u} \cdot \Delta \mathrm{div}\, \boldsymbol{u} ]\, dx \\ &= \displaystyle \int_{\mathbb{R}^3} \left(\left|\mathbf{curl}\, \boldsymbol{u} \right|{}^2 + \left|\mathrm{div} \, \boldsymbol{u}\right|{}^2 \right) \, dx. \end{array} \end{aligned}$$

    Since \(\mathscr {D}(\mathbb {R}^3)^3\) is dense in \(H^1(\mathbb {R}^3)^3\), we deduce that:

    $$\displaystyle \begin{aligned}\forall \, \boldsymbol{u} \in H^1(\mathbb{R}^3)^3, \int_{\mathbb{R}^3} \left|\nabla \boldsymbol{u}\right|{}^2 \, dx = \int_{\mathbb{R}^3} \left(\left|\mathbf{curl}\, \boldsymbol{u} \right|{}^2 + \left|\mathrm{div}\, \boldsymbol{u} \right|{}^2 \right) \, dx. \end{aligned}$$
  3. (iii)

    Back to point (i) of the remark: since \(\boldsymbol {v} \in H^1_0(\Omega )^3\), we have:

    $$\displaystyle \begin{aligned}\left\|\nabla \boldsymbol{v}\right\|{}^2_{L^2(\Omega)} = \left\|\nabla \widetilde{\boldsymbol{v}}\right\|{}_{L^2(\mathbb{R}^3)} = \int_{\mathbb{R}^3} \left( \left|\mathbf{curl}\, \widetilde{\boldsymbol{v}}\right|{}^2 + \left|\mathrm{div}\, \widetilde{\boldsymbol{v}}\right|{}^2 \right)\, dx, \end{aligned}$$

    which gives the relation

    $$\displaystyle \begin{aligned}\int_\Omega \left|\nabla \boldsymbol{v}\right|{}^2 \, dx = \int_\Omega \left(\left|\mathbf{curl}\, \boldsymbol{v}\right|{}^2 + \left|\mathrm{div}\, \boldsymbol{v}\right|{}^2 \right) \, dx. \end{aligned}$$

    Note that this last relation can also be directly established if \(\boldsymbol {v} \in \mathscr {D}(\Omega )^3\) and then, by density of \(\mathscr {D}(\Omega )\) in \(H^1_0(\Omega )^3\), for any \(\boldsymbol {v} \in H^1_0 (\Omega )^3.\)

Remark 16

  1. (i)

    If Ω is simply connected, then for any v ∈ H 1( Ω)3 such that v ⋅ν = 0 on Γ, the inequality (17) is written

    $$\displaystyle \begin{aligned}\left\|\boldsymbol{v}\right\|{}_{H^1(\Omega)^3} \leq C(\Omega) \, \left( \left\|\mathrm{div}\, \boldsymbol{v}\right\|{}_{L^2(\Omega)} + \left\|\mathbf{curl}\, \boldsymbol{v} \right\|{}_{L^2(\Omega)} \right). \end{aligned}$$
  2. (ii)

    If Γ is connected (I = 1), then for any v ∈ H 1( Ω)3 such that v ×ν = 0 on Γ, the inequality (18) is written

    $$\displaystyle \begin{aligned}\left\|\boldsymbol{v}\right\|{}_{H^1(\Omega)^3} \leq C(\Omega) \, \left( \left\|\mathrm{div}\, \boldsymbol{v}\right\|{}_{L^2(\Omega)} + \left\|\mathbf{curl}\, \boldsymbol{v} \right\|{}_{L^2(\Omega)} \right). \end{aligned}$$

Proposition 2.6

Let Ω be a bounded open subset of class \(\mathscr {C}^{1,1}\) of \(\mathbb {R}^3.\) Then the bilinear form

$$\displaystyle \begin{aligned}(\boldsymbol{u},\boldsymbol{v}) \longmapsto \int_{\Omega} \mathbf{curl}\, \boldsymbol{u} \cdot \mathbf{curl}\, \boldsymbol{v} \, dx \end{aligned}$$

is coercive on the following spaces V and on W, respectively:

$$\displaystyle \begin{aligned}V = \left\{\boldsymbol{v} \in H^1(\Omega)^3; \; \mathrm{div}\, \boldsymbol{v} = 0 \; \mathit{\mbox{in}} \; \Omega, \boldsymbol{v} \cdot\boldsymbol{\nu} = 0 \; \mathit{\mbox{on }}\, \Gamma \; \mathit{\mbox{and}} \; \int_{\Sigma_j} \boldsymbol{v} \cdot\boldsymbol{\nu} = 0, 1 \leq j \leq J \right\} \end{aligned}$$
$$\displaystyle \begin{aligned}W = \left\{\boldsymbol{v} \in H^1(\Omega)^3; \mathrm{div}\, \boldsymbol{v} = 0 \; \mathit{\mbox{in}} \; \Omega, \boldsymbol{v} \times\boldsymbol{\nu} = \mathbf{0} \; \mathit{\mbox{on }}\, \Gamma \; \mathit{\mbox{and}} \; \int_{\Gamma_i} \boldsymbol{v} \cdot\boldsymbol{\nu} = 0, 1 \leq j \leq I \right\}. \end{aligned}$$

We are now able to study the problem \((P_{TN}^0)\). We start with the simplest case where Ω is simply connected.

Theorem 2.7

Let Ω be a bounded open domain of class \(\mathscr {C}^{1,1}\) of \(\mathbb {R}^3.\) Suppose that Ω is simply connected.

  1. (i)

    Then for any f ∈ L 6∕5( Ω)3 , Problem \((P_{TN}^0)\) admits a unique solution verifying the estimate

    $$\displaystyle \begin{aligned}\left\|\boldsymbol{u}\right\|{}_{H^1(\Omega)} \leq C(\Omega) \, \left\|\boldsymbol{f}\right\|{}_{L^{6/5}(\Omega)}. \end{aligned}$$
  2. (ii)

    The problem \((P_{TN}^0)\) is equivalent to the problem \((S_{TN}^0)\).

  3. (iii)

    If moreover Ω is of class \(\mathscr {C}^{1, 1}\) then the solution (u, π) ∈ W 2, 6∕5( Ω)3 × W 1, 6∕5( Ω).

Proof

  1. (i)

    The open Ω being simply connected, then

    $$\displaystyle \begin{aligned}V = \left\{\boldsymbol{v} \in H^1(\Omega)^3; \; \mathrm{div}\, \boldsymbol{v} = 0 \quad \mbox{in} \; \Omega,\; \boldsymbol{v} \cdot\boldsymbol{\nu} = 0 \quad \mbox{on }\; \Gamma \right\} \end{aligned}$$

    and V is an Hilbert space. Then let us put

    $$\displaystyle \begin{aligned}a(\boldsymbol{u},\boldsymbol{v}) = \int_\Omega \mathbf{curl}\, \boldsymbol{u} \cdot\mathbf{curl}\, \boldsymbol{v} \, dx. \end{aligned}$$

    Proposition 2.6 shows that the form a is coercive on V . Finally, the form (v) =∫Ω f ⋅v dx is clearly continuous because the continuous embedding H 1( Ω)3L 6( Ω)3. The Lax–Milgram theorem implies the existence of a unique solution of Problem \((P_{TN}^0)\).

  2. (ii)

    Let us first show that

    $$\displaystyle \begin{aligned}(S_{TN}^0) \Longrightarrow (P_{TN}^0). \end{aligned}$$

    Set

    $$\displaystyle \begin{aligned}H = \left\{\boldsymbol{v} \in L^6(\Omega)^3;\; \mathrm{div}\, \boldsymbol{v} \in L^2(\Omega), \boldsymbol{v} \cdot\boldsymbol{\nu} = 0 \; \mbox{on }\; \Gamma \right\}. \end{aligned}$$

    We know that \(\mathscr {D}(\Omega )^3\) is dense in H. So we can show that the dual of H can be characterized as follows:

    $$\displaystyle \begin{aligned}H^\prime = \left\{\boldsymbol{g} + \nabla \chi; \; \boldsymbol{g} \in L^{6/5}(\Omega)^3\; \mathrm{and}\; \chi \in L^2(\Omega) \right\} \end{aligned}$$

    (similar proof to the characterization of the dual H −1( Ω) of \(H^1_0(\Omega )\)).

Let now (u, π) ∈ V × L 2( Ω) solution of \((S_{TN}^0)\). Then for any v ∈ V

$$\displaystyle \begin{aligned}\langle \nabla \pi, \boldsymbol{v} \rangle_{H^\prime \times H} = - \int_\Omega \pi \, \mathrm{div}\, \boldsymbol{v} \, dx = 0. \end{aligned}$$

Therefore,

$$\displaystyle \begin{aligned}- \Delta \boldsymbol{u} = \nabla \pi - \boldsymbol{f}\in H^\prime. \end{aligned}$$

We need the following lemma:

Lemma 2.8

  1. (i)

    The space \(\mathscr {D}(\overline {\Omega })^3\) is dense in the following space

    $$\displaystyle \begin{aligned}E= \left\{\boldsymbol{v} \in H^1(\Omega)^3; \quad \Delta \boldsymbol{v} \in H^\prime \right\}. \end{aligned}$$
  2. (ii)

    The mapping

    $$\displaystyle \begin{aligned}\boldsymbol{v} \longmapsto \mathbf{curl}\, \boldsymbol{v}\times\boldsymbol{\nu} \end{aligned}$$

    defined on \(\mathscr {D}(\overline {\Omega })^3\) can be uniquely extended into a continuous linear mapping from E into H −1∕2( Γ)3.

  3. (iii)

    Moreover, for any φ ∈ H 1( Ω)3 such that

    $$\displaystyle \begin{aligned}div \, \boldsymbol{\varphi} = 0 \; \mathit{\mbox{in}} \; \Omega\quad \mathit{\mbox{and}} \quad \boldsymbol{\varphi} \cdot\boldsymbol{\nu} = 0 \; \mathit{\mbox{on}} \; \Gamma \end{aligned}$$

    and for any v ∈ E, we have the following Green formula

    $$\displaystyle \begin{aligned}- \langle \Delta \boldsymbol{v},\boldsymbol{\varphi} \rangle_{H^\prime \times H} = \int_\Omega \mathbf{curl}\, \boldsymbol{v} \cdot\mathbf{curl}\, \boldsymbol{\varphi} \, dx + \langle \mathbf{curl}\, \boldsymbol{v} \times\boldsymbol{\nu}, \boldsymbol{\varphi} \rangle_\Gamma, \end{aligned}$$

    where 〈⋅, ⋅〉Γ denotes the duality brackets H −1∕2( Γ)3 × H 1∕2( Γ)3.

We return to the proof of the theorem. Since u ∈ H 1( Ω)3 and Δu ∈ H , i.e., u ∈ E, we can use this lemma to deduce on the one hand that the condition curl u = 0 has a meaning in H −1∕2( Γ)3 and, on the other hand, that

$$\displaystyle \begin{aligned}\forall \, \boldsymbol{v} \in V,\quad \langle - \Delta \boldsymbol{u},\boldsymbol{v} \rangle_{H^\prime \times H} = \int_\Omega \mathbf{curl}\, \boldsymbol{u} \cdot \mathbf{curl}\, \boldsymbol{v} \, dx = \int_\Omega \boldsymbol{f}\cdot \boldsymbol{v} \, dx, \end{aligned}$$

i.e., u is solution of \((P_{TN}^0)\).

Conversely, let u ∈ V solution of Problem \((P_{TN}^0)\). Then

$$\displaystyle \begin{aligned}\mathrm{div}\ \, \boldsymbol{u} = 0 \quad \mbox{in} \; \Omega, \quad \boldsymbol{u} \cdot\boldsymbol{\nu} = 0 \; \mbox{on} \; \Gamma \end{aligned}$$

and

$$\displaystyle \begin{aligned}\forall \,\boldsymbol{v} \in \mathscr{D}(\Omega)^3 \; \mbox{with} \; \mathrm{div} \, \boldsymbol{v} = 0 \; \mbox{in} \; \Omega \end{aligned}$$

we have

$$\displaystyle \begin{aligned}\langle \mathbf{curl}\, \mathbf{curl}\, \boldsymbol{u}, \boldsymbol{v} \rangle_{\mathscr{D}^\prime(\Omega)^3 \times \mathscr{D}(\Omega)^3} = \langle \boldsymbol{f},\boldsymbol{v} \rangle_{\mathscr{D}^\prime(\Omega)^3 \times \mathscr{D}(\Omega)^3}. \end{aligned}$$

That gives

$$\displaystyle \begin{aligned}\langle - \Delta \boldsymbol{u}, \boldsymbol{v} \rangle_{\mathscr{D}^\prime(\Omega)^3 \times \mathscr{D}(\Omega)^3} = \langle \boldsymbol{f},\boldsymbol{v} \rangle_{\mathscr{D}^\prime(\Omega)^3 \times \mathscr{D}(\Omega)^3}. \end{aligned}$$

So there exists, by De Rham’s theorem, a function π in L 2( Ω), unique up to an additive constant, such that

$$\displaystyle \begin{aligned} - \Delta \boldsymbol{u} - \boldsymbol{f}= \nabla (-\pi) \quad \mbox{in} \; \Omega \end{aligned} $$
(19)

(note that L 6∕5( Ω)↪H −1( Ω)).

It remains to show that u vérifies:

$$\displaystyle \begin{aligned}\mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu} = \mathbf{0} \quad \mbox{on} \; \Gamma. \end{aligned}$$

For that, from (19) and use the formula of Green of the first lemma, one deduces that

$$\displaystyle \begin{aligned}\forall \, \boldsymbol{v} \in V, \quad \langle - \Delta \boldsymbol{u} + \nabla \pi, \boldsymbol{v} \rangle_{H^\prime \times H} = \int_\Omega \mathbf{curl}\, \boldsymbol{u} \cdot \mathbf{curl}\, \boldsymbol{v} \, dx + \langle \mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu},\boldsymbol{v} \rangle_\Gamma \end{aligned}$$

that is to say that

$$\displaystyle \begin{aligned}\forall \, \boldsymbol{v} \in V,\quad \int_\Omega \mathbf{curl}\, \boldsymbol{u}\cdot \mathbf{curl}\, \boldsymbol{v} \, dx + \langle \mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu},\boldsymbol{v} \rangle_\Gamma = \int_\Omega \boldsymbol{f} \cdot \boldsymbol{v} \, dx. \end{aligned}$$

But u being solution of \((P_{TN}^0)\), then

$$\displaystyle \begin{aligned}\forall \, \boldsymbol{v} \in V, \quad \langle \mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu},\boldsymbol{v} \rangle_\Gamma = 0. \end{aligned}$$

Now let it be μ ∈ H 1∕2( Γ). We know that there exists

$$\displaystyle \begin{aligned}\boldsymbol{w} \in H^1(\Omega)^3,\; \mathrm{div}\, \boldsymbol{w} = 0 \; \mbox{in} \; \Omega, \; \boldsymbol{w} = \boldsymbol{\mu}_\tau \; \mbox{on} \; \Gamma, \end{aligned}$$

where μ τ = μ − (μ ⋅ν)ν the tangential component of μ on Γ. As w ∈ V , we have:

$$\displaystyle \begin{aligned}\langle \mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu},\boldsymbol{\mu} \rangle_\Gamma = \langle \mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu},\boldsymbol{\mu}_\tau \rangle_\Gamma = \langle \mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu},\boldsymbol{w} \rangle_\Gamma =0, \end{aligned}$$

which means that

$$\displaystyle \begin{aligned}\mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu} = \mathbf{0} \quad \mbox{on} \; \Gamma. \end{aligned}$$

(iii) The regularity W 1, 6∕5( Ω) of π is due to the fact that π satisfies:

$$\displaystyle \begin{aligned}\mathrm{div}\, (\nabla \pi - \boldsymbol{f}) = 0 \quad \mathrm{in}\; \Omega \quad \mathrm{and}\quad (\nabla \pi - \boldsymbol{f})\cdot \boldsymbol{\nu} = 0 \quad \mathrm{on}\; \Gamma. \end{aligned}$$

Setting z = curl u, the regularity W 2, 6∕5( Ω)3 of u is a consequence of the following properties:

$$\displaystyle \begin{aligned}\boldsymbol{z}\in L^{6/5}(\Omega)^3,\; \mathrm{div}\, \boldsymbol{z} = 0,\; \mathbf{curl}\, \boldsymbol{z} \in L^{6/5}(\Omega)^3\quad \mathrm{and}\quad \boldsymbol{z} \times\boldsymbol{\nu} = \mathbf{0} \; \; \mathrm{on}\; \Gamma. \end{aligned}$$

Case Ω non Simply Connected

We then show that the kernel:

$$\displaystyle \begin{aligned}K_T (\Omega) = \left\{\boldsymbol{v} \in L^2(\Omega)^3; \; \mathrm{div} \boldsymbol{v} = 0, \; \mathbf{curl}\, \boldsymbol{v} = \mathbf{0} \; \mathrm{in}\, \Omega \;\mbox{and} \; \boldsymbol{v} \cdot\boldsymbol{\nu} = 0 \; \mbox{on} \;\Gamma\right\} \end{aligned}$$

is of finite dimension and that the dimension corresponds to the number of cuts Σj necessary to obtain an open set \(\overset {\circ }{\Omega } = \Omega \setminus \displaystyle \cup _{j=1}^J \Sigma _j\) simply connected.

As a consequence, if

$$\displaystyle \begin{aligned}V = \left\{\boldsymbol{v} \in H^1(\Omega)^3;\quad \mathrm{div}\, \boldsymbol{v} = 0 \;\mbox{in } \; \Omega \;\mbox{and} \; \boldsymbol{v} \cdot\boldsymbol{\nu} = 0 \; \mbox{on} \;\Gamma \right\}, \end{aligned}$$

then, to prove that Problem \((P_{TN}^0)\) admits a solution, it is necessary that f satisfies the following compatibility condition:

$$\displaystyle \begin{aligned}\forall \, \boldsymbol{v} \in K_T(\Omega),\quad \int_\Omega \boldsymbol{f}\cdot \boldsymbol{v} \, dx = 0. \end{aligned}$$

Moreover, if such a solution u exists, it is unique up to an additive element of K T( Ω).

2.3 The Stokes Problem with Navier Boundary Condition

We recall the Navier condition:

$$\displaystyle \begin{aligned}\left[2(\mathbb{D}\boldsymbol{u})\boldsymbol{\nu}\right]_{\tau} + \alpha \boldsymbol{u}_\tau = \mathbf{0} \quad \mbox{on} \; \Gamma, \end{aligned}$$

where

$$\displaystyle \begin{aligned}\mathbb{D}\boldsymbol{u} = \left(\frac{1}{2} \left(\frac{ \partial u_i}{\partial x_j} + \frac{ \partial u_j}{\partial x_i} \right)\right)_{1 \leq i,j\leq3} \end{aligned}$$

is the deformation tensor, α defined on Γ is the friction coefficient and u τ is the tangential component of u. To simplify, we will consider here only the case α = 0.

Note that when div u = 0 in Ω, then \(2 \mathrm {div}\ \mathbb {D}\boldsymbol {u} = \Delta \boldsymbol {u}.\)

Lemma 2.9

If (u, π) ∈ H 1( Ω)3 × L 2( Ω) is such that

$$\displaystyle \begin{aligned}- \Delta \boldsymbol{u} + \nabla \pi \in L^{6/5}(\Omega)^3 \end{aligned}$$

then

$$\displaystyle \begin{aligned}\left[(\mathbb{D}\boldsymbol{u})\boldsymbol{\nu}\right]_\tau \in H^{-1/2}(\Gamma)^3 \end{aligned}$$

and

$$\displaystyle \begin{aligned}for \, any \; \boldsymbol{\varphi} \in H^1(\Omega)^3 \quad \mathit{\mbox{such that}} \quad \mathrm{div}\ \, \boldsymbol{\varphi} = 0\quad \mathit{\mbox{in}} \; \Omega \quad \mathit{\mbox{and}} \quad \boldsymbol{\varphi} \cdot\boldsymbol{\nu} = 0 \quad \mathit{\mbox{on}} \;\Gamma \end{aligned}$$

we have the Green’s formula:

$$\displaystyle \begin{aligned}\int_\Omega \left(- \Delta \boldsymbol{u} + \nabla \pi \right) \cdot \boldsymbol{\varphi} \, dx = 2 \int_\Omega \mathbb{D}\boldsymbol{u} : \mathbb{D} \boldsymbol{\varphi} \, dx - 2 \langle \left[(\mathbb{D}\boldsymbol{u})\boldsymbol{\nu}\right]_\tau,\boldsymbol{\varphi} \rangle_\Gamma, \end{aligned}$$

where 〈⋅, ⋅〉Γ denotes the duality brackets H −1∕2( Γ)3 × H −1∕2( Γ)3.

With this Green’s formula, the Stokes problem can be formulated as:

$$\displaystyle \begin{aligned}(P_N^0) \, \left\{ \begin{array}{ll} \mbox{Find} \; \boldsymbol{u} \in V, \; \mbox{such that for any} \, \, \boldsymbol{\varphi} \in V, \\ \displaystyle 2 \int_\Gamma \mathbb{D}\boldsymbol{u} : \mathbb{D} \boldsymbol{\varphi} \, dx = \int_\Omega \boldsymbol{f} \cdot\boldsymbol{\varphi} \, dx. \end{array} \right. \end{aligned}$$

Set

$$\displaystyle \begin{aligned}a(\boldsymbol{u},\boldsymbol{\varphi} ) = \int_\Omega \mathbb{D}\boldsymbol{u} : \mathbb{D} \boldsymbol{\varphi} \, dx. \end{aligned}$$

When Ω is not axisymmetric, then this form is coercive on V due to Korn’s inequality:

$$\displaystyle \begin{aligned}\left\|\boldsymbol{u}\right\|{}_{H^1(\Omega)} \backsimeq \left\|\mathbb{D}\boldsymbol{u}\right\|{}_{L^2(\Omega)}. \end{aligned}$$

While if Ω is axisymmetric, this is not the case anymore. We must then quotient by some finite dimensional kernel.

Remark 17

In fact, on Γ we have the relation:

$$\displaystyle \begin{aligned}\left[2(\mathbb{D}\boldsymbol{u})\boldsymbol{\nu}\right]_\tau = \mathbf{curl}\, \boldsymbol{u} \times\boldsymbol{\nu} - \Lambda \boldsymbol{u}, \end{aligned}$$

where Λ is an operator of order 0:

$$\displaystyle \begin{aligned}\Lambda \boldsymbol{u} = \displaystyle\sum_{k=1}^2 \left(\boldsymbol{u}_\tau \cdot \frac{\partial\boldsymbol{\nu}}{\partial s_k} \right) \boldsymbol{\tau}_k, \end{aligned}$$

where (τ 1, τ 2) is a base of the tangent plane to Γ at point x and (s 1, s 2) are local coordinates in this tangent plane.

This means that on the questions of regularity, they can be reduced to those concerning the Navier type condition.