This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Partial boundary regularity for the Navier–Stokes equations in irregular domains

Dominic Breit Department of Mathematics, Heriot-Watt University, Edinburgh, EH14 4AS, United Kingdom [email protected] Institute of Mathematics, TU Clausthal, Erzstraße 1, 38678 Clausthal-Zellerfeld, Germany [email protected]
Abstract.

We prove partial regularity of suitable weak solutions to the Navier–Stokes equations at the boundary in irregular domains. In particular, we provide a criterion which yields continuity of the velocity field in a boundary point and obtain solutions which are continuous in a.a. boundary boundary point (their existence is a consequence of a new maximal regularity result for the Stokes equations in domains with minimal regularity). We suppose that we have a Lipschitz boundary with locally small Lipschitz constant which belongs to the fractional Sobolev space W21/p,pW^{2-1/p,p} for some p>154p>\frac{15}{4}. The same result was previously only known under the much stronger assumption of a C2C^{2}-boundary.

2010 Mathematics Subject Classification:
76Nxx; 76N10; 35Q30 ; 35Q84; 82D60

1. Introduction

We consider the motion of a viscous incompressible fluid in a physical body Ω\Omega – a bounded Lipschitz domain in 3\mathbb{R}^{3} during the interval =(0,T)\mathcal{I}=(0,T). The motion of the fluid is governed by the Navier–Stokes equations

ϱ(t𝐮+(𝐮)𝐮)\displaystyle\varrho\big{(}\partial_{t}{\bf u}+({\bf u}\cdot\nabla){\bf u}\big{)} =μΔ𝐮π+𝐟,div𝐮=0,\displaystyle=\mu\Delta{\bf u}-\nabla\pi+{\bf f},\quad\operatorname{div}{\bf u}=0, (1)

in ×Ω\mathcal{I}\times\Omega, where 𝐮:×Ω3{\bf u}:\mathcal{I}\times\Omega\rightarrow\mathbb{R}^{3} is the velocity field and π:×Ω\pi:\mathcal{I}\times\Omega\rightarrow\mathbb{R} the pressure function. The quantity 𝐟:×Ω3{\bf f}:\mathcal{I}\times\Omega\rightarrow\mathbb{R}^{3} is an external forcing, ϱ\varrho is the density and μ\mu the viscosity – two positive constants which will be set to 1 in the following for simplicity.

The existence of weak solutions to (1) has been established in the 1930’s by Leray [16]. The regularity of solutions to (1) is an outstanding open problem which has been attracting mathematicians for decades – still we are far away from a complete understanding (though some remarkable recent progress on the non-uniqueness of weak solutions has been made based on the method of convex integration, cf. [1, 4, 5]). The state of the art today is partial regularity. This means that the velocity field is locally bounded/Hölder continuous outside a negligible set of the space-time cylinder (further regularity properties inside this set can be deduced) with measure zero. Such an analysis has been initiated in a series of papers by Sheffer, see [27][30]. A further milestone is the work by Caffarelli-Kohn-Nirenberg in [6] on suitable weak solutions. These solutions satisfy a form of the energy inequality which is localised in space-time (hereafter called local energy inequality) which reads as

Ω12ζ|𝐮(t)|2dx+0tΩζ|𝐮|2dxdσ0tΩ12(|𝐮|2(tζ+Δζ)+(|𝐮|2+2π)𝐮ζ)dxdσ+0tΩζ𝐟𝐮dxdσ\begin{split}\int_{\Omega}\frac{1}{2}&\zeta\big{|}{\bf u}(t)\big{|}^{2}\,\mathrm{d}x+\int_{0}^{t}\int_{\Omega}\zeta|\nabla{\bf u}|^{2}\,\mathrm{d}x\,\mathrm{d}\sigma\\ &\leq\int_{0}^{t}\int_{\Omega}\frac{1}{2}\Big{(}|{\bf u}|^{2}(\partial_{t}\zeta+\Delta\zeta)+\big{(}|{\bf u}|^{2}+2\pi\big{)}{\bf u}\cdot\nabla\zeta\Big{)}\,\mathrm{d}x\,\mathrm{d}\sigma+\int_{0}^{t}\int_{\Omega}\zeta{\bf f}\cdot{\bf u}\,\mathrm{d}x\,\mathrm{d}\sigma\end{split} (2)

for any ζCc(×Ω)\zeta\in C_{c}^{\infty}(\mathcal{I}\times\Omega) with ζ0\zeta\geq 0. This is a piece of information which has to be included into the definition of a solution as it is otherwise lost in the construction procedure. It is not known (and maybe not even expected) if any weak solution satisfies a local energy inequality (in fact, the same problem appears for the global energy inequality too). As demonstrated in [6], an analysis of the local regularity properties of suitable weak solutions is possible. In particular, a criterion for a solution to (1) is provided, which yields boundedness of the velocity field in a given point in space-time. Some further improvements and simplifications can be found in [9, 15, 18].

The regularity of solutions to (1) at the boundary seems to be less understood. A first results has been achieved by Sheffer in [30] and a more systematic analysis was started by Seregin in [23][25]. Still all these results consider the case of a flat boundary. A flat boundary is the easiest case to consider and one expects that the same results also apply for curved boundaries provided they are sufficiently smooth. Indeed, a corresponding theory for non-flat boundaries of class C2C^{2} has been obtained in [26]. Many applications naturally lead, however, to a boundary of significantly less regularity. This is particularly motivated by problems from fluid-structure interaction, where the boundary is described by the displacement of an elastic structure. The latter is the solution to a partial differential equation on its own and hence only of limited regularity (we comment further on this in Section 6.3).

We aim to prove partial regularity of solutions to (1) at the boundary under minimal assumptions on the regularity of the boundary. Our analysis is based on the concept of boundary suitable weak solutions as in [23, 26]. They satisfy a local energy inequality near a certain part of the boundary (that is, (2) for cut-off functions supported near a part of the boundary), see Definition 2.4. So far, even their existence was not known for boundaries with regularity below C2C^{2}. In order to improve this we prove a new result on the maximal LtrLxpL^{r}_{t}L^{p}_{x}-theory for the unsteady Stokes system under minimal assumptions on the boundary regularity, cf. Theorem 3.1. The main assumption is that Ω\Omega is a Lipschitz domain with locally small Lipschitz constant and that the local coordinates (we make this concept precise in Section 2.4.) belong to the class of Sobolev multipliers on W21/p,pW^{2-1/p,p} – the trace space of W2,pW^{2,p} to which the velocity field belongs. The LtrLxpL^{r}_{t}L^{p}_{x}-theory just described provides a parabolic counterpart of the recent results on the steady Stokes system from [2] and yields the existence of boundary suitable weak solutions to the Navier–Stokes system (1) in irregular domains, cf. Theorem 2.5.

Eventually, we prove a criterion for boundary suitable weak solutions to (1) which implies continuity of the velocity field in a boundary point (see Theorem 2.6) and hence obtain solutions which are continuous in almost any boundary boundary point (see Theorem 2.7). Our main assumption is that the boundary coordinates belong to the class of Sobolev multipliers on

W21p,p(2)for somep>154\displaystyle W^{2-\frac{1}{p},p}(\mathbb{R}^{2})\quad\text{for some}\quad p>\tfrac{15}{4} (3)

with sufficiently small norm. This class includes Lipschitz boundaries with small Lipschitz constant belonging to the class Wσ,p(2)W^{\sigma,p}(\mathbb{R}^{2}) for σ>21/p\sigma>2-1/p, see Remark 2.2.

Our approach uses a flattening of the boundary by means of a transformation 𝚽W2,p(3){\boldsymbol{\Phi}}\in W^{2,p}(\mathbb{R}^{3}) which is an extension of the function φ\varphi describing the boundary locally. This leads to some kind of perturbed Navier–Stokes equations (or perturbed Stokes equations if the convective term is neglected). We provide a regularity theory for the perturbed Stokes system under minimal assumptions on the coefficients resulting from the flattening (this is similar to that of the Stokes system in irregular domains mentioned above). This is used in the partial regularity proof via the blow-up technique. There are various stages in the proof of the blow up lemma (see Lemma 5.1), which require restrictions on the regularity of 𝚽{\boldsymbol{\Phi}} (or that of φ\varphi). The most restrictive one is related to the decay of the pressure 𝔮~\tilde{\mathfrak{q}} of some perturbed Stokes system: We have to show that

τ3(ττ+|𝔮~(𝔮~)τ+|5/3dzdσ)95τ2α\displaystyle\tau^{3}\bigg{(}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{I}_{\tau}}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{B}^{+}_{\tau}}|\tilde{\mathfrak{q}}-(\tilde{\mathfrak{q}})_{\mathcal{B}^{+}_{\tau}}|^{5/3}\,\mathrm{d}z\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}}\lesssim\tau^{2\alpha} (4)

for some α>0\alpha>0 to arrive at a contradiction. Here 𝒬τ+=τ×τ+\mathcal{Q}^{+}_{\tau}=\mathcal{I}_{\tau}\times\mathcal{B}_{\tau}^{+} denotes a parabolic half-cylinder centered at the flat boundary with radius τ>0\tau>0. Estimate (4) can be proved by Poincaré’s inequality if 𝔮~\nabla\tilde{\mathfrak{q}} belongs to the space Lt5/3LxpL^{5/3}_{t}L^{p}_{x} with p>154p>\frac{15}{4}. The term in (4) results from our unusual choice of excess functional related to the integrability Lt5/3Lx5/3L_{t}^{5/3}L_{x}^{5/3} for the pressure. In order to fine-tune the assumptions on pp in (3) (see Section 6.2 for some discussion) we must choose the time-integrability of the pressure as large as possible, where the upper limit is 5/35/3 due to the integrability of the convective term.

Details on the LtrLxpL^{r}_{t}L^{p}_{x}-theory for the perturbed Stokes system can be found in Lemma 4.2 (which also implies a useful Caccioppoli-type inequality in Lemma 4.3). The latter also yields an estimate for the velocity field 𝐮¯\overline{\bf u} of the perturbed Stokes system in the same space which implies its continuity (we expect this to be true also under weaker assumptions, see the discussion in Section 6.2). This is needed similarly to (4) for the decay of the perturbed velocity field. Finally, the transformation of the local energy inequality from the original to the flat geometry requires some assumptions on the boundary, though weaker than those already mentioned (we refer to the estimates in (72)). At first glance it seems that one needs 𝚽W2,{\boldsymbol{\Phi}}\in W^{2,\infty} to control this term. A more careful analysis reveals, however, that it is sufficient if 𝚽{\boldsymbol{\Phi}} belongs to the Sobolev multiplier class on W2,3/2W^{2,3/2}. Although it remains unclear at this stage if (3) is optimal for partial boundary regularity, we believe that it will be very difficult to relax it. We comment further on this in Section 6.2.

2. Preliminaries and results

2.1. Conventions

We write fgf\lesssim g for two non-negative quantities ff and gg if there is a c>0c>0 such that fcgf\leq\,cg. Here cc is a generic constant which does not depend on the crucial quantities and can change from line to line. If necessary we specify particular dependencies. We write fgf\approx g if fgf\lesssim g and gfg\lesssim f. We do not distinguish in the notation for the function spaces between scalar- and vector-valued functions. However, vector-valued functions will usually be denoted in bold case.

2.2. Classical function spaces

Let 𝒪m\mathcal{O}\subset\mathbb{R}^{m}, m1m\geq 1, be open. Function spaces of continuous or α\alpha-Hölder continuous functions, α(0,1)\alpha\in(0,1), are denoted by C(𝒪¯)C(\overline{\mathcal{O}}) or C0,α(𝒪¯)C^{0,\alpha}(\overline{\mathcal{O}}) respectively. Similarly, we write C1(𝒪¯)C^{1}(\overline{\mathcal{O}}) and C1,α(𝒪¯)C^{1,\alpha}(\overline{\mathcal{O}}). We denote as usual by Lp(𝒪)L^{p}(\mathcal{O}) and Wk,p(𝒪)W^{k,p}(\mathcal{O}) for p[1,]p\in[1,\infty] and kk\in\mathbb{N} Lebesgue and Sobolev spaces over 𝒪\mathcal{O}. For a bounded domain 𝒪\mathcal{O} the space Lp(𝒪)L^{p}_{\perp}(\mathcal{O}) denotes the subspace of Lp(𝒪)L^{p}(\mathcal{O}) of functions with zero mean, that is (f)𝒪:=𝒪fdx:=m(𝒪)1𝒪fdx=0(f)_{\mathcal{O}}:=\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{O}}f\,\mathrm{d}x:=\mathcal{L}^{m}(\mathcal{O})^{-1}\int_{\mathcal{O}}f\,\mathrm{d}x=0. We denote by W0k,p(𝒪)W^{k,p}_{0}(\mathcal{O}) the closure of the smooth and compactly supported functions in Wk,p(𝒪)W^{k,p}(\mathcal{O}). This coincides with the functions vanishing m1\mathcal{H}^{m-1} -a.e. on 𝒪\partial\mathcal{O} provided 𝒪\partial\mathcal{O} is sufficiently regular. We also denote by Wk,p(𝒪)W^{-k,p}(\mathcal{O}) the dual of W0k,p(𝒪)W^{k,p}_{0}(\mathcal{O}). Finally, we consider the subspace W0,div1,p(𝒪)W^{1,p}_{0,\operatorname{div}}(\mathcal{O}) of divergence-free vector fields which is defined accordingly. We will use the shorthand notations LxpL^{p}_{x} and Wxk,pW^{k,p}_{x} in the case of 33-dimensional domains and LypL^{p}_{y} and Wyk,pW^{k,p}_{y} for 22-dimensional sets.

For a separable Banach space (X,X)(X,\|\cdot\|_{X}) we denote by Lp(0,T;X)L^{p}(0,T;X) the set of (Bochner-) measurable functions u:(0,T)Xu:(0,T)\rightarrow X such that the mapping tu(t)XLp(0,T)t\mapsto\|u(t)\|_{X}\in L^{p}(0,T). The set C([0,T];X)C([0,T];X) denotes the space of functions u:[0,T]Xu:[0,T]\rightarrow X which are continuous with respect to the norm topology on (X,X)(X,\|\cdot\|_{X}). The space W1,p(0,T;X)W^{1,p}(0,T;X) consists of those functions from Lp(0,T;X)L^{p}(0,T;X) for which the distributional time derivative belongs to Lp(0,T;X)L^{p}(0,T;X) as well. We use the shorthand LtpXL^{p}_{t}X for Lp(0,T;X)L^{p}(0,T;X). For instance, we write LtpWx1,pL^{p}_{t}W^{1,p}_{x} for Lp(0,T;W1,p(𝒪))L^{p}(0,T;W^{1,p}(\mathcal{O})). Similarly, Wtk,pXW^{k,p}_{t}X stands for Wk,p(0,T;X)W^{k,p}(0,T;X).

The space Cα,β([0,T]×𝒪¯)C^{\alpha,\beta}([0,T]\times\overline{\mathcal{O}}) with α,β(0,1]\alpha,\beta\in(0,1] denotes the set of functions being α\alpha-Hölder continuous in t[0,T]t\in[0,T] and β\beta-Hölder continuous in x𝒪¯x\in\overline{\mathcal{O}}.

2.3. Fractional differentiability and Sobolev mulitpliers

For p[1,)p\in[1,\infty) the fractional Sobolev space (Sobolev-Slobodeckij space) with differentiability s>0s>0 with ss\notin\mathbb{N} will be denoted by Ws,p(𝒪)W^{s,p}(\mathcal{O}). For s>0s>0 we write s=s+{s}s=\lfloor s\rfloor+\{s\} with s0\lfloor s\rfloor\in\mathbb{N}_{0} and {s}(0,1)\{s\}\in(0,1). We denote by W0s,p(𝒪)W^{s,p}_{0}(\mathcal{O}) the closure of the smooth and compactly supported functions in W1,p(𝒪)W^{1,p}(\mathcal{O}). For s>1ps>\frac{1}{p} this coincides with the functions vanishing m1\mathcal{H}^{m-1} -a.e. on 𝒪\partial\mathcal{O} provided 𝒪\partial\mathcal{O} is regular enough. We also denote by Ws,p(𝒪)W^{-s,p}(\mathcal{O}) for s>0s>0 the dual of W0s,p(𝒪)W^{s,p}_{0}(\mathcal{O}). Similar to the case of unbroken differentiabilities above we use the shorthand notations Wxs,pW^{s,p}_{x} and Wys,pW^{s,p}_{y}.. We will denote by Bp,qs(m)B^{s}_{p,q}(\mathbb{R}^{m}) the standard Besov spaces on m\mathbb{R}^{m} with differentiability s>0s>0, integrability p[1,]p\in[1,\infty] and fine index q[1,]q\in[1,\infty]. They can be defined (for instance) via Littlewood-Paley decomposition leading to the norm Bp,qs(m)\|\cdot\|_{B^{s}_{p,q}(\mathbb{R}^{m})}. We refer to [21] and [33, 34] for an extensive picture. The Besov spaces Bp,qs(𝒪)B^{s}_{p,q}(\mathcal{O}) for a bounded domain 𝒪m\mathcal{O}\subset\mathbb{R}^{m} are defined as the restriction of functions from Bp,qs(m)B^{s}_{p,q}(\mathbb{R}^{m}), that is

Bp,qs(𝒪)\displaystyle B^{s}_{p,q}(\mathcal{O}) :={f|𝒪:fBp,qs(m)},\displaystyle:=\{f|_{\mathcal{O}}:\,f\in B^{s}_{p,q}(\mathbb{R}^{m})\},
gBp,qs(𝒪)\displaystyle\|g\|_{B^{s}_{p,q}(\mathcal{O})} :=inf{fBp,qs(m):f|𝒪=g}.\displaystyle:=\inf\{\|f\|_{B^{s}_{p,q}(\mathbb{R}^{m})}:\,f|_{\mathcal{O}}=g\}.

If ss\notin\mathbb{N} and p(1,)p\in(1,\infty) we have Bp,ps(𝒪)=Ws,p(𝒪)B^{s}_{p,p}(\mathcal{O})=W^{s,p}(\mathcal{O}).

In accordance with [19, Chapter 14] the Sobolev multiplier norm is given by

φs,p(𝒪):=sup𝐯:𝐯Ws1,p(𝒪)=1φ𝐯Ws1,p(𝒪),\displaystyle\|\varphi\|_{\mathcal{M}^{s,p}(\mathcal{O})}:=\sup_{{\bf v}:\,\|{\bf v}\|_{W^{s-1,p}(\mathcal{O})}=1}\|\nabla\varphi\cdot{\bf v}\|_{W^{s-1,p}(\mathcal{O})}, (5)

where p[1,]p\in[1,\infty] and s1s\geq 1. The space s,p(𝒪)\mathcal{M}^{s,p}(\mathcal{O}) of Sobolev multipliers is defined as those objects for which the s,p(𝒪)\mathcal{M}^{s,p}(\mathcal{O})-norm is finite. For δ>0\delta>0 we denote by s,p(𝒪)(δ)\mathcal{M}^{s,p}(\mathcal{O})(\delta) the subset of functions from s,p(𝒪)\mathcal{M}^{s,p}(\mathcal{O}) with s,p(𝒪)\mathcal{M}^{s,p}(\mathcal{O})-norm not exceeding δ\delta. By mathematical induction with respect to ss one can prove for Lipschitz-continuous functions φ\varphi that membership to s,p(𝒪)\mathcal{M}^{s,p}(\mathcal{O}) in the sense of (5) implies that

supw:wWs,p(𝒪)=1φwWs,p(𝒪)<.\displaystyle\sup_{w:\,\|w\|_{W^{s,p}(\mathcal{O})}=1}\|\varphi\,w\|_{W^{s,p}(\mathcal{O})}<\infty. (6)

The quantity (6) also serves as customary definition of the Sobolev multiplier norm in the literature but (5) is more suitable for our purposes.

Let us finally collect some some useful properties of Sobolev multipliers. By [19, Corollary 14.6.2] we have

ϕs,p(m)ϕL(m),\displaystyle\|\phi\|_{\mathcal{M}^{s,p}(\mathbb{R}^{m})}\lesssim\|\nabla\phi\|_{L^{\infty}(\mathbb{R}^{m})}, (7)

provided that one of the following conditions holds:

  • p(s1)<mp(s-1)<m and ϕBϱ,ps(m)\phi\in B^{s}_{\varrho,p}(\mathbb{R}^{m}) with ϱ[pmp(s1)1,]\varrho\in\big{[}\frac{pm}{p(s-1)-1},\infty\big{]};

  • p(s1)=mp(s-1)=m and ϕBϱ,ps(m)\phi\in B^{s}_{\varrho,p}(\mathbb{R}^{m}) with ϱ(p,]\varrho\in(p,\infty].

Note that the hidden constant in (7) depends on the Bϱ,ps(m)B^{s}_{\varrho,p}(\mathbb{R}^{m})-norm of ϕ\phi. By [19, Corollary 4.3.8] it holds

ϕs,p(m)ϕWs1,p(m)\displaystyle\|\phi\|_{\mathcal{M}^{s,p}(\mathbb{R}^{m})}\approx\|\nabla\phi\|_{W^{s-1,p}(\mathbb{R}^{m})} (8)

for p(s1)>mp(s-1)>m. Finally, we note the following rule about the composition with Sobolev multipliers which is a consequence of [19, Lemma 9.4.1]. For open sets 𝒪1,𝒪2m\mathcal{O}_{1},\mathcal{O}_{2}\subset\mathbb{R}^{m}, uWs,p(𝒪2)u\in W^{s,p}(\mathcal{O}_{2}) and a Lipschitz continuous function ϕ:𝒪1𝒪2{\boldsymbol{\phi}}:\mathcal{O}_{1}\rightarrow\mathcal{O}_{2} with ϕs,p(𝒪1){\boldsymbol{\phi}}\in\mathcal{M}^{s,p}(\mathcal{O}_{1}) and Lipschitz continuous inverse ϕ1:𝒪2𝒪1{\boldsymbol{\phi}}^{-1}:\mathcal{O}_{2}\rightarrow\mathcal{O}_{1} we have

uϕWs,p(𝒪1)uWs,p(𝒪2)\displaystyle\|u\circ{\boldsymbol{\phi}}\|_{W^{s,p}(\mathcal{O}_{1})}\lesssim\|u\|_{W^{s,p}(\mathcal{O}_{2})} (9)

with constant depending on ϕ{\boldsymbol{\phi}}.

2.4. Parametrisation of domains

In this section we present the necessary framework to parametrise the boundary of the underlying domain Ω3\Omega\subset\mathbb{R}^{3} by local maps of a certain regularity. This yields, in particular, a rigorous definition of a s,p\mathcal{M}^{s,p}-boundary. We follow the presentation from [2].

We assume that Ω\partial\Omega can be covered by a finite number of open sets 𝒰1,,𝒰\mathcal{U}^{1},\dots,\mathcal{U}^{\ell} for some \ell\in\mathbb{N}, such that the following holds. For each j{1,,}j\in\{1,\dots,\ell\} there is a reference point yj3y^{j}\in\mathbb{R}^{3} and a local coordinate system {e1j,e2j,e3j}\{e^{j}_{1},e^{j}_{2},e_{3}^{j}\} (which we assume to be orthonormal and set 𝒬j=(e1j|e2j|e3j)3×3\mathcal{Q}_{j}=(e_{1}^{j}|e_{2}^{j}|e_{3}^{j})\in\mathbb{R}^{3\times 3}), a function φj:2\varphi_{j}:\mathbb{R}^{2}\rightarrow\mathbb{R} and rj>0r_{j}>0 with the following properties:

  1. (A1)

    There is hj>0h_{j}>0 such that

    𝒰j={x=𝒬jz+yj3:z=(z,z3)3,|z|<rj,|z3φj(z)|<hj}.\mathcal{U}^{j}=\{x=\mathcal{Q}_{j}z+y^{j}\in\mathbb{R}^{3}:\,z=(z^{\prime},z_{3})\in\mathbb{R}^{3},\,|z^{\prime}|<r_{j},\,|z_{3}-\varphi_{j}(z^{\prime})|<h_{j}\}.
  2. (A2)

    For x𝒰jx\in\mathcal{U}^{j} we have with z=𝒬jt(xyj)z=\mathcal{Q}_{j}^{t}(x-y^{j})

    • xΩx\in\partial\Omega if and only if z3=φj(z)z_{3}=\varphi_{j}(z^{\prime});

    • xΩx\in\Omega if and only if 0<z3φj(z)<hj0<z_{3}-\varphi_{j}(z^{\prime})<h_{j};

    • xΩx\notin\Omega if and only if 0>z3φj(z)>hj0>z_{3}-\varphi_{j}(z^{\prime})>-h_{j}.

  3. (A3)

    We have that

    Ωj=1𝒰j.\partial\Omega\subset\bigcup_{j=1}^{\ell}\mathcal{U}^{j}.

In other words, for any x0Ωx_{0}\in\partial\Omega there is a neighborhood UU of x0x_{0} and a function ϕ:2\phi:\mathbb{R}^{2}\rightarrow\mathbb{R} such that after translation and rotation111By translation via yjy_{j} and rotation via 𝒬j\mathcal{Q}_{j} we can assume that x0=0x_{0}=0 and that the outer normal at x0x_{0} is pointing in the negative x3x_{3}-direction.

UΩ=UG,G={(x,x3)3:x2,x3>ϕ(x)}\displaystyle U\cap\Omega=U\cap G,\quad G={\{{(x^{\prime},x_{3})\in\mathbb{R}^{3}\,:\,x^{\prime}\in\mathbb{R}^{2},x_{3}>\phi(x^{\prime})}\}} (10)

The regularity of Ω\partial\Omega will be described by means of local coordinates as just described.

Definition 2.1.

Let Ω3\Omega\subset\mathbb{R}^{3} be a bounded domain, s1s\geq 1 and p[1,]p\in[1,\infty]. We say that Ω\partial\Omega belongs to the class s,p\mathcal{M}^{s,p} if there is \ell\in\mathbb{N} and functions φ1,,φs,p(2)\varphi_{1},\dots,\varphi_{\ell}\in\mathcal{M}^{s,p}(\mathbb{R}^{2}) satisfying (A1)(A3).

Clearly, we can define similarly a s,p(δ)\mathcal{M}^{s,p}(\delta)-boundary for some δ>0\delta>0 by requiring that φ1,,φs,p(2)(δ)\varphi_{1},\dots,\varphi_{\ell}\in\mathcal{M}^{s,p}(\mathbb{R}^{2})(\delta). Analogous definitions apply for various other function spaces such as Bϱ,psB^{s}_{\varrho,p} for s>0s>0 and ϱ,p[1,]\varrho,p\in[1,\infty] or C1,αC^{1,\alpha} for α(0,1]\alpha\in(0,1]. Of particular importance for us is also a Lipschitz boundary, where φ1,,φW1,(2)\varphi_{1},\dots,\varphi_{\ell}\in W^{1,\infty}(\mathbb{R}^{2}). We say that the Lipschitz constant of Ω\partial\Omega, denoted by Lip(Ω)\mathrm{Lip}(\partial\Omega), is (smaller or) equal to some number L>0L>0 provided the Lipschitz constants of φ1,,φ\varphi_{1},\dots,\varphi_{\ell} are not exceeding LL.

Remark 2.2.

It follows from (7) and (8) that Ωs,p(δ)\partial\Omega\in\mathcal{M}^{s,p}(\delta) provided Ω\Omega is a Lipschitz domain with sufficiently small Lipschitz constant and ΩBϱ,pθ\partial\Omega\in B^{\theta}_{\varrho,p} for θ>s\theta>s and

ϱpifp(s1)3,ϱ2pp(s1)1ifp(s1)<3,\displaystyle\varrho\geq p\quad\text{if}\quad p(s-1)\geq 3,\quad\varrho\geq\tfrac{2p}{p(s-1)-1}\quad\text{if}\quad p(s-1)<3, (11)

The Lipschitz constant can be made sufficiently small, for instance, if ΩC1,α\partial\Omega\in C^{1,\alpha} for some α>0\alpha>0.

In order to describe the behaviour of functions defined in Ω\Omega close to the boundary we need to extend the functions φ1,,φ\varphi_{1},\dots,\varphi_{\ell} from (A1)(A3) to the half space :={ξ=(ξ,ξ3):ξ3>0}\mathbb{H}:={\{{\xi=(\xi^{\prime},\xi_{3})\,:\,\xi_{3}>0}\}}. Hence we are confronted with the task of extending a function ϕ:2\phi\,:\,\mathbb{R}^{2}\to\mathbb{R} to a mapping Φ:3\Phi\,:\,\mathbb{H}\to\mathbb{R}^{3} that maps the 0-neighborhood in \mathbb{H} to the x0x_{0}-neighborhood in Ω\Omega. The mapping (ξ,0)(ξ,ϕ(ξ))(\xi^{\prime},0)\mapsto(\xi^{\prime},\phi(\xi^{\prime})) locally maps the boundary of \mathbb{H} to the one of Ω\partial\Omega. We extend this mapping using the extension operator of Maz’ya and Shaposhnikova [19, Section 9.4.3]. Let ζCc(B1(0))\zeta\in C^{\infty}_{c}(B_{1}(0^{\prime})) with ζ0\zeta\geq 0 and 2ζ(x)dx=1\int_{\mathbb{R}^{2}}\zeta(x^{\prime})\,\mathrm{d}x^{\prime}=1. Let ζt(x):=t2ζ(x/t)\zeta_{t}(x^{\prime}):=t^{-2}\zeta(x^{\prime}/t) denote the induced family of mollifiers. We define the extension operator

(𝒯ϕ)(ξ,ξ3)=2ζξ3(ξy)ϕ(y)dy,(ξ,ξ3),\displaystyle(\mathcal{T}\phi)(\xi^{\prime},\xi_{3})=\int_{\mathbb{R}^{2}}\zeta_{\xi_{3}}(\xi^{\prime}-y^{\prime})\phi(y^{\prime})\,\mathrm{d}y^{\prime},\quad(\xi^{\prime},\xi_{3})\in\mathbb{H},

where ϕ:2\phi:\mathbb{R}^{2}\to\mathbb{R} is a Lipschitz function with Lipschitz constant LL. Then the estimate

(𝒯ϕ)Ws,p(3)cϕWs1p,p(2)\displaystyle{\lVert{\nabla(\mathcal{T}\phi)}\rVert}_{W^{s,p}(\mathbb{R}^{3})}\leq c{\lVert{\nabla\phi}\rVert}_{W^{s-\frac{1}{p},p}(\mathbb{R}^{2})} (12)

follows from [19, Theorem 8.7.2]. Moreover, [19, Theorem 8.7.1] yields

𝒯ϕs,p()ϕs1p,p(2).\displaystyle\|\mathcal{T}\phi\|_{\mathcal{M}^{s,p}(\mathbb{H})}\lesssim\|\phi\|_{\mathcal{M}^{s-\frac{1}{p},p}(\mathbb{R}^{2})}. (13)

It is shown in [19, Lemma 9.4.5] that (for sufficiently large NN, i.e., Nc(ζ)L+1N\geq c(\zeta)L+1) the mapping

αz(z3)Nz3+(𝒯ϕ)(z,z3)\displaystyle\alpha_{z^{\prime}}(z_{3})\mapsto N\,z_{3}+(\mathcal{T}\phi)(z^{\prime},z_{3})

is for every z2z^{\prime}\in\mathbb{R}^{2} one to one and the inverse is Lipschitz with gradient bounded by (NL)1(N-L)^{-1}. Now, we define the mapping 𝚽:3{\boldsymbol{\Phi}}\,:\,\mathbb{H}\to\mathbb{R}^{3} as a rescaled version of the latter one by setting

𝚽(ξ,ξ3)\displaystyle{\boldsymbol{\Phi}}(\xi^{\prime},\xi_{3}) :=(ξ,αξ3(ξ))=(ξ,ξ3+(𝒯ϕ)(ξ,ξ3/N)).\displaystyle:=\big{(}\xi^{\prime},\alpha_{\xi_{3}}(\xi^{\prime})\big{)}=\big{(}\xi^{\prime},\,\xi_{3}+(\mathcal{T}\phi)(\xi^{\prime},\xi_{3}/N)\big{)}. (14)

Thus, 𝚽{\boldsymbol{\Phi}} is one-to-one (for sufficiently large N=N(L)N=N(L)) and we can define its inverse 𝚿:=𝚽1{\boldsymbol{\Psi}}:={\boldsymbol{\Phi}}^{-1}. The Jacobi matrix of the mapping 𝚽{\boldsymbol{\Phi}} satisfies

J=𝚽=(𝕀2×20ξ(𝒯ϕ)1+1/Nξn𝒯ϕ).\displaystyle J=\nabla{\boldsymbol{\Phi}}=\begin{pmatrix}\mathbb{I}_{2\times 2}&0\\ \partial_{\xi^{\prime}}(\mathcal{T}\phi)&1+1/N\partial_{\xi_{n}}\mathcal{T}\phi\end{pmatrix}. (15)

Since |ξ3𝒯ϕ|L{\lvert{\partial_{\xi_{3}}\mathcal{T}\phi}\rvert}\leq L, we have

12<1L/N|det(J)|1+L/N2\displaystyle\frac{1}{2}<1-L/N\leq{\lvert{\det(J)}\rvert}\leq 1+L/N\leq 2 (16)

using that NN is large compared to LL. Finally, we note the implication

𝚽s,p()𝚿s,p(),\displaystyle{\boldsymbol{\Phi}}\in\mathcal{M}^{s,p}(\mathbb{H})\,\,\Rightarrow\,\,{\boldsymbol{\Psi}}\in\mathcal{M}^{s,p}(\mathbb{H}), (17)

which holds, for instance, if 𝚽{\boldsymbol{\Phi}} is Lipschitz continuous, cf. [19, Lemma 9.4.2].

Remark 2.3.
  1. (a)

    Since the cover 𝒰1,,𝒰\mathcal{U}^{1},\dots,\mathcal{U}^{\ell} is open it is possible to find a number >0\mathfrak{R}>0 (depending on the cover) such that the following holds: for every xΩx\in\partial\Omega there is j{1,,}j\in\{1,\dots,\ell\} such that x𝒰jx\in\mathcal{U}^{j} and dist(x,3𝒰j)\mathrm{dist}(x,\mathbb{R}^{3}\setminus\mathcal{U}^{j})\geq\mathfrak{R}.

  2. (b)

    Similarly, by possibly decreasing \mathfrak{R}, we have the following: there is δ>0\delta>0 (depending on the cover) such that for any xΩx\in\Omega with dist(x,Ω)δ\mathrm{dist}(x,\partial\Omega)\leq\delta there is j{1,,}j\in\{1,\dots,\ell\} such that x𝒰jx\in\mathcal{U}^{j} and dist(x,3𝒰j)\mathrm{dist}(x,\mathbb{R}^{3}\setminus\mathcal{U}^{j})\geq\mathfrak{R}.

2.5. The main results

We start with a definition of boundary suitable weak solutions adapting the notation from [26]. These solutions satisfy a local form of the energy inequality in the neighborhood of boundary points. For that purpose we fix two numbers r(1,2)r_{\ast}\in(1,2) and s(1,32)s_{\ast}\in(1,\frac{3}{2}) such that 1r+32s2\frac{1}{r_{\ast}}+\frac{3}{2s_{\ast}}\geq 2. The choice comes from the fact that the convective term (𝐮)𝐮(\nabla{\bf u}){\bf u} of a weak solution to (1) belongs to LtrLxsL^{r_{\ast}}_{t}L^{s_{\ast}}_{x}. Later on we will choose r=5/3r_{\ast}=5/3 and, accordingly, s=15/14s_{\ast}=15/14.

Definition 2.4 (Boundary suitable weak solution).

Let Ω3\Omega\subset\mathbb{R}^{3} be a bounded Lipschitz domain and ΓΩ\Gamma\subset\partial\Omega relatively open. Let (𝐟,𝐮0)({\bf f},{\bf u}_{0}) be a dataset such that

𝐟Lr(;Ls(Ω)),𝐮0W2,sW0,div1,2(Ω).\displaystyle{\bf f}\in L^{r_{\ast}}\big{(}\mathcal{I};L^{s_{\ast}}(\Omega)\big{)},\quad{\bf u}_{0}\in W^{2,s_{\ast}}\cap W^{1,2}_{0,\mathrm{\operatorname{div}}}(\Omega). (18)

We call the triple (𝐮,π,φ)({\bf u},\pi,\varphi) a boundary suitable weak solution to the Navier–Stokes system (1) hear Γ\Gamma with data (𝐟,𝐮0)({\bf f},{\bf u}_{0}) provided that the following holds:

  • (a)

    The velocity field 𝐮{\bf u} satisfies

    𝐮L(;L2(Ω))L2(;W0,div1,2(Ω))Lr(;W2,s(Ω))W1,r(;Ls(Ω)).\displaystyle{\bf u}\in L^{\infty}\big{(}\mathcal{I};L^{2}(\Omega)\big{)}\cap L^{2}\big{(}\mathcal{I};W^{1,2}_{0,\operatorname{div}}(\Omega)\big{)}\cap L^{r_{\ast}}(\mathcal{I};W^{2,s_{\ast}}(\Omega))\cap W^{1,r_{\ast}}(\mathcal{I};L^{s_{\ast}}(\Omega)).
  • (b)

    The pressure π\pi satisfies

    πLr(;W1,s(Ω)).\pi\in L^{r_{\ast}}(\mathcal{I};W^{1,s_{\ast}}_{\perp}(\Omega)).
  • c)

    We have222after translation via yjy_{j} and rotation via 𝒬j\mathcal{Q}_{j}, cf. Section 2.4.

    ΓΩG,G={(x,x3)n:x2,x3=φ(x)}\displaystyle\Gamma\subset\partial\Omega\cap G,\quad G={\{{(x^{\prime},x_{3})\in\mathbb{R}^{n}\,:\,x^{\prime}\in\mathbb{R}^{2},x_{3}=\varphi(x^{\prime})}\}}

    for some Lipschitz function φ:2\varphi:\mathbb{R}^{2}\rightarrow\mathbb{R} satisfying

    φ(0)=0,φ(0)=0.\displaystyle\varphi(0)=0,\quad\nabla\varphi(0)=0. (19)
  • (d)

    We have

    t𝐮+(𝐮)𝐮\displaystyle\partial_{t}{\bf u}+({\bf u}\cdot\nabla){\bf u} =Δ𝐮π+𝐟,div𝐮=0,𝐮|Ω=0,𝐮(0,)=𝐮0,\displaystyle=\Delta{\bf u}-\nabla\pi+{\bf f},\quad\operatorname{div}{\bf u}=0,\quad{\bf u}|_{\partial\Omega}=0,\quad{\bf u}(0,\cdot)={\bf u}_{0},

    a.a. in ×Ω\mathcal{I}\times\Omega.

  • (e)

    for any ζCc(×3)\zeta\in C_{c}^{\infty}(\mathcal{I}\times\mathbb{R}^{3}) with ζ0\zeta\geq 0 and spt(ζ)×(3Γ)\mathrm{spt}(\zeta)\subset\mathcal{I}\times(\mathbb{R}^{3}\setminus\Gamma) the local energy inequality

    Ω12ζ|𝐮(t)|2dx+0tΩζ|𝐮|2dxdσ0tΩ12(|𝐮|2(tζ+Δζ)+(|𝐮|2+2π)𝐮ζ)dxdσ+0tΩζ𝐟𝐮dxdσ.\begin{split}\int_{\Omega}\frac{1}{2}&\zeta\big{|}{\bf u}(t)\big{|}^{2}\,\mathrm{d}x+\int_{0}^{t}\int_{\Omega}\zeta|\nabla{\bf u}|^{2}\,\mathrm{d}x\,\mathrm{d}\sigma\\ &\leq\int_{0}^{t}\int_{\Omega}\frac{1}{2}\Big{(}|{\bf u}|^{2}(\partial_{t}\zeta+\Delta\zeta)+\big{(}|{\bf u}|^{2}+2\pi\big{)}{\bf u}\cdot\nabla\zeta\Big{)}\,\mathrm{d}x\,\mathrm{d}\sigma+\int_{0}^{t}\int_{\Omega}\zeta{\bf f}\cdot{\bf u}\,\mathrm{d}x\,\mathrm{d}\sigma.\end{split} (20)

    holds.

The next theorem shows that, under suitable assumptions on Ω\partial\Omega, there is a solution to (1) which is a boundary suitable weak solution around every boundary point.

Theorem 2.5.

Suppose that Ω3\Omega\subset\mathbb{R}^{3} is a bounded Lipschitz domain such that Lip(Ω)δ\mathrm{Lip}(\partial\Omega)\leq\delta and Ω21/s,s(δ)\partial\Omega\in\mathcal{M}^{2-1/s_{\ast},s_{\ast}}(\delta) for some sufficiently small δ\delta. Then there is a solution (𝐮,π)({\bf u},\pi) to the Navier–Stokes equations (1) with the following property: For every point x0Ωx_{0}\in\partial\Omega there is a neighborhood 𝒰(x0)\mathcal{U}(x_{0}) such that (𝐮,π)({\bf u},\pi) generates a boundary suitable weak solution to the Navier–Stokes system (1) near 𝒰(x0)Ω\mathcal{U}(x_{0})\cap\partial\Omega in the sense of Definition 2.4.

Proof.

Applying a standard regularisation procedure (by convolution with a mollifying kernel) to the functions φ1,,φ\varphi_{1},\dots,\varphi_{\ell} from (A1)(A3) in the parametrisation of Ω\partial\Omega we obtain a smooth boundary. Classically, the solution to the corresponding Stokes system is smooth. Such a procedure is standard and has been applied, for instance, in [7, 8]. It is possible to do this in a way that the original domain is included in the regularised domain to which we extend the functions 𝐟{\bf f} and 𝐮0{\bf u}_{0} by means of an extension operator. The regularisation applied to the φjs\varphi_{j}^{\prime}s converges on all Besov spaces with p<p<\infty. As shown in [19, Lemma 4.3.3.] it does not expand the s,p(2)\mathcal{M}^{s,p}(\mathbb{R}^{2})-norm, which is sufficient. For a smooth domain the statement of Theorem 2.5 is well-known (it can, for instance, be proved along the lines of [6]). We obtain a sequence of functions (𝐮m,πm)({\bf u}_{m},\pi_{m}) which satisfy the Navier–Stokes equations as well as the local energy inequality (20). Clearly, they can be constructed to also satisfy the global energy inequality. Hence our sequence (𝐮m)({\bf u}_{m}) is bounded in the energy space

L(;L2(Ω))L2(;W0,div1,2(Ω)).\displaystyle L^{\infty}(\mathcal{I};L^{2}(\Omega))\cap L^{2}(\mathcal{I};W^{1,2}_{0,\operatorname{div}}(\Omega)).

As a consequence, we can bound the convective term (𝐮m)𝐮m(\nabla{\bf u}_{m}){\bf u}_{m} in Lr(I;Ls(Ω))L^{r_{\ast}}(I;L^{s_{\ast}}(\Omega)). Now we come to the crucial point: By the maximal regularity theory for the Stokes system from Theorem 3.1 below we have

t𝐮mLr(;Ls(Ω))+𝐮mLr(;W2,s(Ω))+πmLr(;W1,s(Ω))𝐟Lr(;Ls(Ω)+(𝐮m)𝐮mLr(;Ls(Ω)+𝐮0W2,s(Ω)\displaystyle\begin{aligned} \|\partial_{t}{\bf u}_{m}\|_{L^{r_{\ast}}(\mathcal{I};L^{s_{\ast}}(\Omega))}&+\|{\bf u}_{m}\|_{L^{r_{\ast}}(\mathcal{I};W^{2,s_{\ast}}(\Omega))}+\|\pi_{m}\|_{L^{r_{\ast}}(\mathcal{I};W^{1,s_{\ast}}(\Omega))}\\ &\lesssim\|{\bf f}\|_{L^{r_{\ast}}(\mathcal{I};L^{s_{\ast}}(\Omega)}+\|(\nabla{\bf u}_{m}){\bf u}_{m}\|_{L^{r_{\ast}}(\mathcal{I};L^{s_{\ast}}(\Omega)}+\|{\bf u}_{0}\|_{W^{2,s_{\ast}}(\Omega)}\end{aligned} (21)

uniformly in mm. With (21) at hand, we obtain (after passing to a subsequence) limit objects with the claimed regularity and can pass to the limit in the momentum equation and local energy inequality. ∎

Theorem 2.6.

Let Ω3\Omega\subset\mathbb{R}^{3} be a bounded Lipschitz domain and ΓΩ\Gamma\subset\partial\Omega relatively open. Suppose that 𝐟Lp(;Lp(Ω)){\bf f}\in L^{p}(\mathcal{I};L^{p}(\Omega)) for some p>154p>\frac{15}{4}. There is a number ε0>0\varepsilon_{0}>0 such that the following holds. Let (𝐮,π,φ)({\bf u},\pi,\varphi) be a boundary suitable weak solution to the Navier–Stokes system (1) near Γ\Gamma in the sense of Definition 2.4, where φ21/p,p(2)(δ)\varphi\in\mathcal{M}^{2-1/p,p}(\mathbb{R}^{2})(\delta) with φWy1,δ\|\varphi\|_{W^{1,\infty}_{y}}\leq\delta and sufficiently small δ>0\delta>0. Let x0Γx_{0}\in\Gamma and t0It_{0}\in I such that

r2t0r2t0+r2Ωr(x0)|𝐮|3dxdt+(r5/3t0r2t0+r2Ωr(x0)|π|5/3dxdt)95<ε0\displaystyle r^{-2}\int_{t_{0}-r^{2}}^{t_{0}+r^{2}}\int_{\Omega\cap\mathcal{B}_{r}(x_{0})}|{\bf u}|^{3}\,\mathrm{d}x\,\mathrm{d}t+\bigg{(}r^{-5/3}\int_{t_{0}-r^{2}}^{t_{0}+r^{2}}\int_{\Omega\cap\mathcal{B}_{r}(x_{0})}|\pi|^{5/3}\,\mathrm{d}x\,\mathrm{d}t\bigg{)}^{\frac{9}{5}}<\varepsilon_{0} (22)

for some rdis(x0,3Γ)r\leq\mathrm{dis}(x_{0},\mathbb{R}^{3}\setminus\Gamma). Then we have 𝐮C0,α(𝒰¯(t0,x0)){\bf u}\in C^{0,\alpha}(\overline{\mathcal{U}}(t_{0},x_{0})) for some α>0\alpha>0 and a neighborhood 𝒰(t0,x0)\mathcal{U}(t_{0},x_{0}) of (t0,x0)(t_{0},x_{0}).

Denoting by paras\mathcal{H}^{s}_{\mathrm{para}} the ss-dimensional parabolic Hausdorff measure (and using its definition based on covering with parabolic cubes) it is standard to deduce the following result concerning the size of the singular set from Theorem 2.6.

Theorem 2.7.

Suppose that Ω3\Omega\subset\mathbb{R}^{3} is a bounded Lipschitz domain such that Lip(Ω)δ\mathrm{Lip}(\partial\Omega)\leq\delta and Ω21/p,p(δ)\partial\Omega\in\mathcal{M}^{2-1/p,p}(\delta) for some p>154p>\frac{15}{4} and sufficiently small δ\delta. Suppose that 𝐟Lp(;Lp(Ω)){\bf f}\in L^{p}(\mathcal{I};L^{p}(\Omega)). Then there is a solution (𝐮,π)({\bf u},\pi) to the Navier–Stokes equations (1) and a closed set ΣI×Ω\Sigma\subset I\times\partial\Omega with para5/3(Σ)=0\mathcal{H}^{5/3}_{\mathrm{para}}(\Sigma)=0 such that for any (t0,x0)I×ΩΣ(t_{0},x_{0})\in I\times\partial\Omega\setminus\Sigma we have 𝐮C0,α(𝒰¯(t0,x0)){\bf u}\in C^{0,\alpha}(\overline{\mathcal{U}}(t_{0},x_{0})) for some α>0\alpha>0 and a neighborhood 𝒰(t0,x0)\mathcal{U}(t_{0},x_{0}) of (t0,x0)(t_{0},x_{0}).

Our result in Theorem 2.7 is in terms of the size of the singular set weaker than the result in [26] for more regular domains. It is shown there that the dimension of the singular set is one rather than 5/35/3. We comment on this gap in more detail in Section 6.1. It is unlcear if this is an intrinsic feature of irregular domains or a drawback of our method.

3. The Stokes system in irregular domains

In this section we consider the unsteady Stokes system

t𝐮=Δ𝐮π+𝐟,div𝐮=0,𝐮|I×Ω=𝐮,𝐮(0,)=𝐮0,\displaystyle\partial_{t}{\bf u}=\Delta{\bf u}-\nabla\pi+{\bf f},\quad\operatorname{div}{\bf u}=0,\quad{\bf u}|_{I\times\partial{\Omega}}={\bf u}_{\partial},\quad{\bf u}(0,\cdot)={\bf u}_{0}, (23)

in a domain Ω3{\Omega}\subset\mathbb{R}^{3} with unit normal 𝐧{\bf n}. The result given in the following theorem is a maximal regularity estimate for the solution in terms of the right-hand side under minimal assumption on the regularity of Ω\partial{\Omega} (see Remark 11 for the connecton between Sobolev multipliers and Besov spaces).

Theorem 3.1.

Let p,r(1,)p,r\in(1,\infty) and suppose that Ω{\Omega} is a Lipschitz domain with local Lipschitz constant δ\delta belonging to the class 21/p,p(2)(δ)\mathcal{M}^{2-1/p,p}(\mathbb{R}^{2})(\delta) for some sufficiently small δ\delta, 𝐟Lr(;Lp(Ω)){\bf f}\in L^{r}(\mathcal{I};L^{p}({\Omega})) and 𝐮0W2,p(Ω)Ldiv2(Ω){\bf u}_{0}\in W^{2,p}({\Omega})\cap L^{2}_{\operatorname{div}}({\Omega}) with tr𝐮0=𝐮\mathrm{tr}\,{\bf u}_{0}={\bf u}_{\partial}, where 𝐮W21/p,p(Ω){\bf u}_{\partial}\in W^{2-1/p,p}(\partial{\Omega}) with Ω𝐮𝐧d2=0\int_{\partial{\Omega}}{\bf u}_{\partial}\cdot{\bf n}\,\mathrm{d}\mathcal{H}^{2}=0. Then there is a unique solution to (23) and we have

t𝐮Lr(;Lp(Ω))+𝐮Lr(;W2,p(Ω))+πLr(;W1,p(Ω))𝐟Lr(;Lp(Ω))+𝐮0W2,p(Ω)+𝐮W21/p,p(Ω).\displaystyle\begin{aligned} \|\partial_{t}{\bf u}\|_{L^{r}(\mathcal{I};L^{p}({\Omega}))}&+\|{\bf u}\|_{L^{r}(\mathcal{I};W^{2,p}({\Omega}))}+\|\pi\|_{L^{r}(\mathcal{I};W^{1,p}({\Omega}))}\\ &\lesssim\|{\bf f}\|_{L^{r}(\mathcal{I};L^{p}({\Omega}))}+\|{\bf u}_{0}\|_{W^{2,p}({\Omega})}+\|{\bf u}_{\partial}\|_{W^{2-1/p,p}(\partial{\Omega})}.\end{aligned} (24)
Remark 3.2.

As shown in [19, Chapter 14] the assumptions on Ω\partial\Omega in Theorem 3.1 are sharp already for the Laplace equation.

Remark 3.3.

Theorem 3.1 provides a parabolic counterpart of the result in the steady Stokes system from [2, Theorem 3.1] (with the same assumptions on Ω\partial\Omega). A direct adaption of the ideas used in the proof of the steady analogue from [2, Theorem 3.1] is not straightforward due to the appearance of the time-derivative in the divergence-correction. Hence we follow instead the classical approach from [31] (see also the presentation in [3, Appendix 2]). The idea is to first solve the problem in the flat geometry and built the solution of the original problem by concatenating the (transformed) solutions. This is somewhat the opposite way compared to [2, Theorem 3.1] and leads to various lower order error terms. They can be controlled for small times and we obtain a global-in-time solution by gluing local solutions together.

Remark 3.4.

Since 𝒪\mathcal{O} is assumed to be a Lipschitz domain the Bogovskii operator has the usual properties, cf. (25) below, and we can extend the result of Theorem 3.1 to the Stokes problem with a given divergence hLr(;W1,p(Ω))W1,r(;W1,p(𝒪))h\in L^{r}(\mathcal{I};W^{1,p}({\Omega}))\cap W^{1,r}(\mathcal{I};W^{-1,p}(\mathcal{O})). In this case the additional terms

hLr(;W1,p(𝒪)),hW1,r(;W1,p(𝒪)),\|h\|_{L^{r}(\mathcal{I};W^{1,p}(\mathcal{O}))},\quad\|h\|_{W^{1,r}(\mathcal{I};W^{-1,p}(\mathcal{O}))},

appear on the right-hand side of (48).

Proof.

By use of a standard extension operator we can assume that 𝐮=0{\bf u}_{\partial}=0. Otherwise, we can solve the homogeneous problem with solution 𝐮~\tilde{{\bf u}} and set

𝐮:=𝐮~+Ω𝐮BogΩ(divΩ𝐮){\bf u}:=\tilde{\bf u}+\mathcal{E}_{{\Omega}}{\bf u}_{\partial}-\operatorname{Bog}_{\Omega}(\operatorname{div}\mathcal{E}_{{\Omega}}{\bf u}_{\partial})

where

Ω:W21/p,p(Ω)W2,p(Ω)\mathcal{E}_{{\Omega}}:W^{2-1/p,p}(\partial{\Omega})\rightarrow W^{2,p}({\Omega})

is a continuous linear extension operator and BogΩ\operatorname{Bog}_{\Omega} the Bogovskii-operator. The latter solves the divergence equation (with respect to homogeneous boundary conditions on Ω\partial{\Omega}) and satisfies

BogΩdiv:W2,p{𝐰:Ω𝐰𝐧d2=0}W02,p(Ω)\displaystyle\operatorname{Bog}_{\Omega}\operatorname{div}:W^{2,p}\cap\bigg{\{}{\bf w}:\,\int_{\partial{\Omega}}{\bf w}\cdot{\bf n}\,\mathrm{d}\mathcal{H}^{2}=0\bigg{\}}\rightarrow W^{2,p}_{0}({\Omega}) (25)

for all p(1,)p\in(1,\infty), see [13][Section III.3].

We want to invert the operator

\displaystyle\mathscr{L} :𝒴r,pLr(;Ldivp(𝒪)),𝐯𝒫p(t𝐯Δ𝐯),\displaystyle:\mathscr{Y}_{r,p}\rightarrow L^{r}(\mathcal{I};L^{p}_{\operatorname{div}}(\mathcal{O})),\quad{\bf v}\mapsto\mathcal{P}_{p}\big{(}\partial_{t}{\bf v}-\Delta{\bf v}\big{)},

where the space 𝒴r,p\mathscr{Y}_{r,p} is given by

𝒴r,p:=Lr(;W0,div1,pW2,p(Ω))W1,r(;Lp(Ω)){𝐯(0,)=0}\displaystyle\mathscr{Y}_{r,p}:=L^{r}(\mathcal{I};W^{1,p}_{0,\operatorname{div}}\cap W^{2,p}({\Omega}))\cap W^{1,r}(\mathcal{I};L^{p}({\Omega}))\cap{\{{{\bf v}(0,\cdot)=0}\}}

and 𝒫p\mathcal{P}_{p} is the Helmholtz projection from Lp(Ω)L^{p}({\Omega}) onto Ldivp(Ω)L^{p}_{\operatorname{div}}({\Omega}). The Helmholtz-projection 𝒫p𝐮\mathcal{P}_{p}{\bf u} of a function 𝐮Lp(Ω){\bf u}\in L^{p}({\Omega}) is defined as 𝒫p𝐮:=𝐮h\mathcal{P}_{p}{\bf u}:={\bf u}-\nabla h, where hh is the solution to the Neumann-problem

{Δh=div𝐮inΩ,𝐧(h𝐮)=0onΩ.\displaystyle\begin{cases}\Delta h=\operatorname{div}{\bf u}\quad\text{in}\quad{\Omega},\\ {\bf n}\cdot(\nabla h-{\bf u})=0\quad\text{on}\quad\partial{\Omega}.\end{cases}

We will try to find an operator :Lr(;Ldivp(Ω))𝒴r,p\mathscr{R}:L^{r}(\mathcal{I};L^{p}_{\operatorname{div}}(\Omega))\rightarrow\mathscr{Y}_{r,p} such that

=id+𝒯,\displaystyle\mathscr{L}\circ\mathscr{R}=\mathrm{id}+\mathscr{T}, (26)

where the operator-norm of 𝒯\mathscr{T} is strictly smaller than 1. This implies that the range of \mathscr{L}\circ\mathscr{R} (which then equals to Lr(;Ldivp(Ω))L^{r}(\mathcal{I};L^{p}_{\operatorname{div}}(\Omega))) is contained in the range of \mathscr{L}. Hence \mathscr{L} is onto.

By assumption there is \ell\in\mathbb{N} and functions φ1,,φ21/p,p(δ)(2)\varphi_{1},\dots,\varphi_{\ell}\in\mathcal{M}^{2-1/p,p}(\delta)(\mathbb{R}^{2}) satisfying (A1)(A3). We clearly find an open set 𝒰0Ω\mathcal{U}^{0}\Subset{\Omega} such that Ωj=0𝒰j{\Omega}\subset\cup_{j=0}^{\ell}\mathcal{U}^{j}. Finally, we consider a decomposition of unity (ξj)j=0(\xi_{j})_{j=0}^{\ell} with respect to the covering 𝒰0,,𝒰\mathcal{U}^{0},\dots,\mathcal{U}^{\ell} of Ω{\Omega}. For j{1,,}j\in\{1,\dots,\ell\} we consider the extension 𝚽j{\boldsymbol{\Phi}}_{j} of φj\varphi_{j} given by (12) with inverse 𝚿j{\boldsymbol{\Psi}}_{j}. Denoting by 𝒱jx=𝒬j(xyj)\mathscr{V}_{j}x=\mathcal{Q}_{j}^{\top}(x-y_{j}) (with the translation yjy_{j} and the rotation 𝒬j\mathcal{Q}_{j} used in (A1)(A3), cf. Section 2.4) we define the operators333Note that by means of a standard extension operator we extend functions in (27)–(29) to the whole space or half space when necessary.

0𝐟\displaystyle\mathscr{R}_{0}{\bf f} :=ξ0𝐔0+j=1ξj𝐔j𝚿j𝒱j,𝒫𝐟:=j=1ξj𝔮j𝚿j𝒱j.\displaystyle:=\xi_{0}{\bf U}_{0}+\sum_{j=1}^{\ell}\xi_{j}{\bf U}_{j}\circ{\boldsymbol{\Psi}}_{j}\circ\mathscr{V}_{j},\quad\mathscr{P}{\bf f}:=\sum_{j=1}^{\ell}\xi_{j}\mathfrak{q}_{j}\circ{\boldsymbol{\Psi}}_{j}\circ\mathscr{V}_{j}. (27)

Here the functions (𝐔0,𝔮0)({\bf U}_{0},\mathfrak{q}_{0}) and (𝐔j,𝔮j)({\bf U}_{j},\mathfrak{q}_{j}) for j{1,,}j\in\{1,\dots,\ell\} are the solutions to the Stokes problem on the whole space and the half space with data 𝐟{\bf f} respectively (transformed if necessary), that is, we have

t𝐔0=Δ𝐔0𝔮0+𝐟,div𝐔0=0,𝐔0|I×=0,𝐔0(0,)=0,\displaystyle\partial_{t}{\bf U}_{0}=\Delta{\bf U}_{0}-\nabla\mathfrak{q}_{0}+{\bf f},\quad\operatorname{div}{\bf U}_{0}=0,\quad{\bf U}_{0}|_{I\times\partial\mathbb{H}}=0,\quad{\bf U}_{0}(0,\cdot)=0, (28)

and

t𝐔j=Δ𝐔j𝔮j+𝐟𝒱j1𝚽j,div𝐔j=0,𝐔j|I×=0,𝐔j(0,)=0.\displaystyle\partial_{t}{\bf U}_{j}=\Delta{\bf U}_{j}-\nabla\mathfrak{q}_{j}+{\bf f}\circ\mathscr{V}_{j}^{-1}\circ{\boldsymbol{\Phi}}_{j},\quad\operatorname{div}{\bf U}_{j}=0,\quad{\bf U}_{j}|_{I\times\partial\mathbb{H}}=0,\quad{\bf U}_{j}(0,\cdot)=0. (29)

We have

(t𝐔0Lxpr+2𝐔0Lxpr+𝔮0Lxpr)dt𝐟Lxprdt,(t𝐔0Wx1,pr+𝐔0Wx1,pr+𝔮0Lxpr)dtTr/2𝐟Lxprdt,\displaystyle\begin{aligned} \int_{\mathcal{I}}\Big{(}\|\partial_{t}{\bf U}_{0}\|_{L^{p}_{x}}^{r}+\|\nabla^{2}{\bf U}_{0}\|^{r}_{L^{p}_{x}}+\|\nabla\mathfrak{q}_{0}\|_{L^{p}_{x}}^{r}\Big{)}\,\mathrm{d}t&\lesssim\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t,\\ \int_{\mathcal{I}}\Big{(}\|\partial_{t}{\bf U}_{0}\|_{W^{-1,p}_{x}}^{r}+\|{\bf U}_{0}\|_{W^{1,p}_{x}}^{r}+\|\mathfrak{q}_{0}\|_{L^{p}_{x}}^{r}\Big{)}\,\mathrm{d}t&\lesssim T^{r/2}\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t,\end{aligned} (30)

and for j=1,,j=1,\dots,\ell

(t𝐔jLxpr+2𝐔jLxpr+𝔮jLxrr)dt𝐟𝒱j𝚽jLxprdt,(t𝐔jWx1,pr+𝐔jWx1,pr+𝔮jLxpr)dtTr/2𝐟𝒱j𝚽jLxprdt,\displaystyle\begin{aligned} \int_{\mathcal{I}}\Big{(}\|\partial_{t}{\bf U}_{j}\|_{L^{p}_{x}}^{r}+\|\nabla^{2}{\bf U}_{j}\|_{L^{p}_{x}}^{r}+\|\nabla\mathfrak{q}_{j}\|_{L^{r}_{x}}^{r}\Big{)}\,\mathrm{d}t&\lesssim\int_{\mathcal{I}}\|{\bf f}\circ\mathscr{V}_{j}\circ{\boldsymbol{\Phi}}_{j}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t,\\ \int_{\mathcal{I}}\Big{(}\|\partial_{t}{\bf U}_{j}\|_{W^{-1,p}_{x}}^{r}+\|{\bf U}_{j}\|_{W^{1,p}_{x}}^{r}+\|\mathfrak{q}_{j}\|_{L^{p}_{x}}^{r}\Big{)}\,\mathrm{d}t&\lesssim T^{r/2}\int_{\mathcal{I}}\|{\bf f}\circ\mathscr{V}_{j}\circ{\boldsymbol{\Phi}}_{j}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t,\end{aligned} (31)

uniformly in TT. Note that estimates (30)2 and (31)2 only hold locally in space (that is, in balls BnB\subset\mathbb{R}^{n} with a constant depending on the radius). This does not cause any problems due to the localisation functions appearing in (27). Since Lipschitz continuity of φj\varphi_{j} implies that of 𝚽j{\boldsymbol{\Phi}}_{j}, cf. (15), we can control the right-hand sides in (31) by 𝐟LtrLxpr\|{\bf f}\|_{L^{r}_{t}L^{p}_{x}}^{r}. Estimates (30) and (31) are classical in the case r=pr=p, see [31, Theorems 3.1 & 3.2]. For the case of arbitrary exponents pp and rr we refer to [14] and the references therein. Note that the TT-dependence in (30)2 and (31)2 follows by simple scaling argument.

The divergence of 0𝐟\mathscr{R}_{0}{\bf f} as defined in (27) is in general not zero. This will be corrected by setting

𝐟=0𝐟+1𝐟,1𝐟=BogΩdiv0𝐟\displaystyle\mathscr{R}{\bf f}=\mathscr{R}_{0}{\bf f}+\mathscr{R}_{1}{\bf f},\quad\mathscr{R}_{1}{\bf f}=-\operatorname{Bog}_{{\Omega}}\operatorname{div}\mathscr{R}_{0}{\bf f}

with the Bogovskii-operator BogΩ\operatorname{Bog}_{\Omega}, cf. (25). Now we clearly have 𝐟𝒴r,p\mathscr{R}{\bf f}\in\mathscr{Y}_{r,p} and the aim is to establish (26). Transforming 𝐔j{\bf U}_{j} and 𝔮j\mathfrak{q}_{j} back to Ω{\Omega}, that is, setting 𝐕j=𝐔j𝚿j𝒱j{\bf V}_{j}={\bf U}_{j}\circ{\boldsymbol{\Psi}}_{j}\circ\mathscr{V}_{j} and 𝔔j=𝔮j𝚿j𝒱j\mathfrak{Q}_{j}=\mathfrak{q}_{j}\circ{\boldsymbol{\Psi}}_{j}\circ\mathscr{V}_{j}, we obtain

t𝐕j=Δ𝐕j𝔔j+(1det(𝚿j))t𝐕jdiv((𝕀3×3𝐀j)𝐕j)div((𝐁j𝕀3×3)𝔔j)+𝐟,div𝐕j=(𝕀3×3𝐁j):𝐕j,𝐕j|Ω𝒰j=0,𝐕j(0,)=0,\displaystyle\begin{aligned} \partial_{t}{\bf V}_{j}=&\Delta{\bf V}_{j}-\nabla\mathfrak{Q}_{j}+(1-\mathrm{det}(\nabla{\boldsymbol{\Psi}}_{j}))\partial_{t}{\bf V}_{j}-\operatorname{div}\big{(}(\mathbb{I}_{3\times 3}-{\bf A}_{j})\nabla{\bf V}_{j})-\operatorname{div}((\mathbf{B}_{j}-\mathbb{I}_{3\times 3})\mathfrak{Q}_{j})+{\bf f},\\ &\operatorname{div}{\bf V}_{j}=(\mathbb{I}_{3\times 3}-\mathbf{B}_{j})^{\top}:\nabla{\bf V}_{j},\quad{\bf V}_{j}|_{\partial{\Omega}\cap\mathcal{U}^{j}}=0,\quad{\bf V}_{j}(0,\cdot)=0,\end{aligned} (32)

where 𝐀j:=det(𝚿j)𝚽j𝚿j𝚽j𝚿j{\bf A}_{j}:=\mathrm{det}(\nabla{\boldsymbol{\Psi}}_{j})\nabla{\boldsymbol{\Phi}}_{j}^{\top}\circ{\boldsymbol{\Psi}}_{j}\nabla{\boldsymbol{\Phi}}_{j}\circ{\boldsymbol{\Psi}}_{j} and 𝐁j:=det(𝚿j)𝚽j𝚿j\mathbf{B}_{j}:=\mathrm{det}(\nabla{\boldsymbol{\Psi}}_{j})\nabla{\boldsymbol{\Phi}}_{j}\circ{\boldsymbol{\Psi}}_{j}. There holds

t𝐟\displaystyle\partial_{t}\mathscr{R}{\bf f} Δ𝐟+𝒫𝐟=𝐟+𝒮𝐟+(tΔ)1𝐟,\displaystyle-\Delta\mathscr{R}{\bf f}+\nabla\mathscr{P}{\bf f}={\bf f}+\mathscr{S}{\bf f}+(\partial_{t}-\Delta)\mathscr{R}_{1}{\bf f}, (33)
𝒮𝐟\displaystyle\mathscr{S}{\bf f} =𝐕0ξ0div(ξ0𝐕0)j=1𝐕jξj\displaystyle=-\nabla{\bf V}_{0}\nabla\xi_{0}-\operatorname{div}\big{(}\nabla\xi_{0}\otimes{\bf V}_{0}\big{)}-\sum_{j=1}^{\ell}\nabla{\bf V}_{j}\nabla\xi_{j}
j=1div(ξj𝐕j)+j=1ξj𝔔jj=1ξjdiv((𝐁j𝕀3×3)𝔔j)\displaystyle-\sum_{j=1}^{\ell}\operatorname{div}\big{(}\nabla\xi_{j}\otimes{\bf V}_{j}\big{)}+\sum_{j=1}^{\ell}\nabla\xi_{j}\mathfrak{Q}_{j}-\sum_{j=1}^{\ell}\xi_{j}\operatorname{div}((\mathbf{B}_{j}-\mathbb{I}_{3\times 3})\mathfrak{Q}_{j}) (34)
j=1ξjdiv((𝕀3×3𝐀j)𝐕j)+j=1ξj(1det(𝚿j))t𝐕j=:i=18𝒮i𝐟.\displaystyle-\sum_{j=1}^{\ell}\xi_{j}\operatorname{div}\big{(}(\mathbb{I}_{3\times 3}-{\bf A}_{j})\nabla{\bf V}_{j})+\sum_{j=1}^{\ell}\xi_{j}(1-\mathrm{det}(\nabla{\boldsymbol{\Psi}}_{j}))\partial_{t}{\bf V}_{j}=:\sum_{i=1}^{8}\mathscr{S}_{i}{\bf f}.

From (33) it follows

𝐟=𝐟+𝒫p𝒮𝐟+𝒫p(tΔ)1𝐟,\displaystyle\mathscr{L}\mathscr{R}{\bf f}={\bf f}+\mathcal{P}_{p}\mathscr{S}{\bf f}+\mathcal{P}_{p}(\partial_{t}-\Delta)\mathscr{R}_{1}{\bf f},

i.e., (26) with 𝒯=𝒫p𝒮+𝒫p(tΔ)1\mathscr{T}=\mathcal{P}_{p}\mathscr{S}+\mathcal{P}_{p}(\partial_{t}-\Delta)\mathscr{R}_{1}. We claim that

i=15𝒮i𝐟Lxprdtδ(T)𝐟Lxprdt\displaystyle\sum_{i=1}^{5}\int_{\mathcal{I}}\|\mathscr{S}_{i}{\bf f}\|^{r}_{L^{p}_{x}}\,\mathrm{d}t\leq\,\delta(T)\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t (35)

with δ(T)0\delta(T)\rightarrow 0 for T0T\rightarrow 0.444Note that this will also depend on \ell which we have to choose sufficiently large to obtain maxjφjWy1,Lip(Ω)\max_{j}\|\varphi_{j}\|_{W^{1,\infty}_{y}}\approx\mathrm{Lip}(\partial{\Omega}). Estimate (35) follows from estimates (30) and (31). We can translate (31) into an estimate for 𝐕j{\bf V}_{j} and 𝔔j\mathfrak{Q}_{j} via

𝐕jWxσ,p𝐔jWxσ,p\displaystyle\begin{aligned} \|{\bf V}_{j}\|_{W^{\sigma,p}_{x}}\lesssim\|{\bf U}_{j}\|_{W^{\sigma,p}_{x}}\end{aligned} (36)

for σ{1,2}\sigma\in\{1,2\} and similarly

𝔔jWxσ1,p𝔮jWxσ1,p\displaystyle\begin{aligned} \|\mathfrak{Q}_{j}\|_{W^{\sigma-1,p}_{x}}\lesssim\|\mathfrak{q}_{j}\|_{W^{\sigma-1,p}_{x}}\end{aligned} (37)

recalling estimates (16) and (9). Note that by our assumptions on φj\varphi_{j} and (14) we have 𝚽j2,p(){\boldsymbol{\Phi}}_{j}\in\mathcal{M}^{2,p}(\mathbb{H}) and thus 𝚿j2,p(){\boldsymbol{\Psi}}_{j}\in\mathcal{M}^{2,p}(\mathbb{H}) by (17). Combining the previous arguments proves (35).

Now we are concerned with 𝒮6\mathscr{S}_{6} and 𝒮7\mathscr{S}_{7} obtaining by (15), (9) and the definitions of 𝐀j{\bf A}_{j} and 𝚽j{\boldsymbol{\Phi}}_{j}

ξjdiv\displaystyle\|\xi_{j}\operatorname{div} ((𝕀3×3𝐀j)𝐕j)Lp()\displaystyle\big{(}(\mathbb{I}_{3\times 3}-{\bf A}_{j})\nabla{\bf V}_{j})\|_{L^{p}(\mathbb{H})}
sup𝐰Wx1,p1(𝕀3×3𝐀j)𝐰W1,p()𝐕jWx1,p\displaystyle\lesssim\sup_{\|{\bf w}\|_{W^{1,p}_{x}}\leq 1}\|(\mathbb{I}_{3\times 3}-{\bf A}_{j}){\bf w}\|_{W^{1,p}(\mathbb{H})}\|\nabla{\bf V}_{j}\|_{W^{1,p}_{x}}
sup𝐰W1,p()1(1det(𝚿j))𝐰W1,p()𝐕jWx2,p\displaystyle\lesssim\sup_{\|{\bf w}\|_{W^{1,p}(\mathbb{H})}\leq 1}\|(1-\mathrm{det}(\nabla{\boldsymbol{\Psi}}_{j})){\bf w}\|_{W^{1,p}(\mathbb{H})}\|{\bf V}_{j}\|_{W^{2,p}_{x}}
+sup𝐰Wx1,p1det(𝚿j)(𝕀3×3𝚽j𝚿j)𝐰W1,p()𝐕jWx2,p\displaystyle+\sup_{\|{\bf w}\|_{W^{1,p}_{x}}\leq 1}\|\mathrm{det}(\nabla{\boldsymbol{\Psi}}_{j})(\mathbb{I}_{3\times 3}-\nabla{\boldsymbol{\Phi}}_{j}^{\top}\circ{\boldsymbol{\Psi}}_{j}){\bf w}\|_{W^{1,p}(\mathbb{H})}\|{\bf V}_{j}\|_{W^{2,p}_{x}}
+sup𝐰Wx1,p1det(𝚿j)𝚽j𝚿j(𝕀3×3𝚽j𝚿j)𝐰W1,p()𝐕jWx2,p\displaystyle+\sup_{\|{\bf w}\|_{W^{1,p}_{x}}\leq 1}\|\mathrm{det}(\nabla{\boldsymbol{\Psi}}_{j})\nabla{\boldsymbol{\Phi}}_{j}^{\top}\circ{\boldsymbol{\Psi}}_{j}(\mathbb{I}_{3\times 3}-\nabla{\boldsymbol{\Phi}}_{j}\circ{\boldsymbol{\Psi}}_{j}){\bf w}\|_{W^{1,p}(\mathbb{H})}\|{\bf V}_{j}\|_{W^{2,p}_{x}}
𝒯ϕj2,p()𝐯Wx2,p\displaystyle\lesssim\|\mathcal{T}\phi_{j}\|_{\mathcal{M}^{2,p}(\mathbb{H})}\|{\bf v}\|_{W^{2,p}_{x}}
+𝚿j2,p()3sup𝐰Wx1,p1(𝕀3×3𝚽j𝚿j)𝐰W1,p()𝐕jWx2,p\displaystyle+\|{\boldsymbol{\Psi}}_{j}\|_{\mathcal{M}^{2,p}(\mathbb{H})}^{3}\sup_{\|{\bf w}\|_{W^{1,p}_{x}}\leq 1}\|(\mathbb{I}_{3\times 3}-\nabla{\boldsymbol{\Phi}}_{j}\circ{\boldsymbol{\Psi}}_{j}){\bf w}\|_{W^{1,p}(\mathbb{H})}\|{\bf V}_{j}\|_{W^{2,p}_{x}}
+𝚿j2,p()3𝚽j2,p()sup𝐰Wx1,p1(𝕀3×3𝚽j𝚿j)𝐰W1,p()𝐕jWx2,p\displaystyle+\|{\boldsymbol{\Psi}}_{j}\|_{\mathcal{M}^{2,p}(\mathbb{H})}^{3}\|{\boldsymbol{\Phi}}_{j}\|_{\mathcal{M}^{2,p}(\mathbb{H})}\sup_{\|{\bf w}\|_{W^{1,p}_{x}}\leq 1}\|(\mathbb{I}_{3\times 3}-\nabla{\boldsymbol{\Phi}}_{j}\circ{\boldsymbol{\Psi}}_{j}){\bf w}\|_{W^{1,p}(\mathbb{H})}\|{\bf V}_{j}\|_{W^{2,p}_{x}}
(𝒯ϕj2,p()+sup𝐰Wx1,p1(𝕀3×3𝚽j𝚿j)𝐰W1,p())𝐕jWx2,p,\displaystyle\lesssim\Big{(}\|\mathcal{T}\phi_{j}\|_{\mathcal{M}^{2,p}(\mathbb{H})}+\sup_{\|{\bf w}\|_{W^{1,p}_{x}}\leq 1}\|(\mathbb{I}_{3\times 3}-\nabla{\boldsymbol{\Phi}}_{j}\circ{\boldsymbol{\Psi}}_{j}){\bf w}\|_{W^{1,p}(\mathbb{H})}\Big{)}\|{\bf V}_{j}\|_{W^{2,p}_{x}},

where

sup𝐰Wx1,p1\displaystyle\sup_{\|{\bf w}\|_{W^{1,p}_{x}}\leq 1} (𝕀3×3𝚽j𝚿j)𝐰W1,p()\displaystyle\|(\mathbb{I}_{3\times 3}-\nabla{\boldsymbol{\Phi}}_{j}\circ{\boldsymbol{\Psi}}_{j}){\bf w}\|_{W^{1,p}(\mathbb{H})}
𝒯ϕj𝚿j2,p()𝒯ϕj2,p().\displaystyle\lesssim\|\mathcal{T}\phi_{j}\circ{\boldsymbol{\Psi}}_{j}\|_{\mathcal{M}^{2,p}(\mathbb{H})}\lesssim\|\mathcal{T}\phi_{j}\|_{\mathcal{M}^{2,p}(\mathbb{H})}.

So we finally have

ξjdiv((𝕀3×3𝐀j)𝐕j)Lp()\displaystyle\|\xi_{j}\operatorname{div}\big{(}(\mathbb{I}_{3\times 3}-{\bf A}_{j})\nabla{\bf V}_{j})\|_{L^{p}(\mathbb{H})} 𝒯ϕj2,p()𝐕jW2,p()\displaystyle\lesssim\|\mathcal{T}\phi_{j}\|_{\mathcal{M}^{2,p}(\mathbb{H})}\|{\bf V}_{j}\|_{W^{2,p}(\mathbb{H})}

and, similarly,

ξjdiv((𝐁j𝕀3×3)𝔔j)Lp()\displaystyle\|\xi_{j}\operatorname{div}((\mathbf{B}_{j}-\mathbb{I}_{3\times 3})\mathfrak{Q}_{j})\|_{L^{p}(\mathbb{H})} sup𝐰W1,p()1(𝐁j𝕀3×3)𝐰W1,p()𝔔jW1,p()\displaystyle\lesssim\sup_{\|{\bf w}\|_{W^{1,p}(\mathbb{H})}\leq 1}\|(\mathbf{B}_{j}-\mathbb{I}_{3\times 3}){\bf w}\|_{W^{1,p}(\mathbb{H})}\|\mathfrak{Q}_{j}\|_{W^{1,p}(\mathbb{H})}
𝒯ϕj(W2,p())𝔔jW1,p().\displaystyle\lesssim\|\mathcal{T}\phi_{j}\|_{\mathcal{M}(W^{2,p}(\mathbb{H}))}\|\mathfrak{Q}_{j}\|_{W^{1,p}(\mathbb{H})}.

By (13) we have

𝒯ϕj2,p()φj21/p,p(),\displaystyle\|\mathcal{T}\phi_{j}\|_{\mathcal{M}^{2,p}(\mathbb{H})}\lesssim\|\varphi_{j}\|_{\mathcal{M}^{2-1/p,p}(\mathbb{H})}, (38)

where the right-hand side is conveniently by assumption. Hence we have

(𝒮6𝐟Lxpr+𝒮7𝐟Lxpr)dtδ(Lip(Ω))j=1(𝐕jWx2,pr+𝔔jWx1,pr)dt.\displaystyle\int_{\mathcal{I}}\Big{(}\|\mathscr{S}_{6}{\bf f}\|^{r}_{L^{p}_{x}}+\|\mathscr{S}_{7}{\bf f}\|_{L^{p}_{x}}^{r}\Big{)}\,\mathrm{d}t\leq\,\delta(\mathrm{Lip}({\Omega}))\sum_{j=1}^{\ell}\int_{\mathcal{I}}\big{(}\|{\bf V}_{j}\|_{W^{2,p}_{x}}^{r}+\|\mathfrak{Q}_{j}\|_{W^{1,p}_{x}}^{r}\big{)}\,\mathrm{d}t.

Using again the estimates for the problem on the half space from (31) as well as (36) and (37) we conclude

(𝒮6𝐟Lxpr+𝒮7𝐟Lxpr)dtδ(Lip(Ω))𝐟Lxprdt.\displaystyle\int_{\mathcal{I}}\Big{(}\|\mathscr{S}_{6}{\bf f}\|^{r}_{L^{p}_{x}}+\|\mathscr{S}_{7}{\bf f}\|^{r}_{L^{p}_{x}}\Big{)}\,\mathrm{d}t\leq\,\delta^{\prime}(\mathrm{Lip}({\Omega}))\int_{\mathcal{I}}\|{\bf f}\|^{r}_{L^{p}_{x}}\,\mathrm{d}t.

Both δ(Lip(Ω))\delta(\mathrm{Lip}({\Omega})) and δ(Lip(Ω))\delta^{\prime}(\mathrm{Lip}({\Omega})) can be chosen conveniently small in dependence on Lip(Ω)\mathrm{Lip}({\Omega}). Similarly, we have

𝒮8𝐟Lxprdtδ(Lip(Ω))j=1t𝐕jLxprdtδ(Lip(Ω))𝐟Lxprdt\displaystyle\int_{\mathcal{I}}\|\mathscr{S}_{8}{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t\leq\,\delta(\mathrm{Lip}({\Omega}))\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\partial_{t}{\bf V}_{j}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t\leq\delta^{\prime}(\mathrm{Lip}({\Omega}))\int_{\mathcal{I}}\|{\bf f}\|^{r}_{L^{p}_{x}}\,\mathrm{d}t

using once more (31) in the last step. In conclusion, choosing first \ell large enough and then TT small enough we can infer that

𝒮𝐟Lxprdt14𝐟Lxprdt.\displaystyle\int_{\mathcal{I}}\|\mathscr{S}{\bf f}\|^{r}_{L^{p}_{x}}\,\mathrm{d}t\leq\,\tfrac{1}{4}\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t. (39)

Now we are going to show the same for (tΔ)1(\partial_{t}-\Delta)\mathscr{R}_{1}. We have

div0𝐟=ξ0𝐔0+j=1ξj𝐕j+j=1ξj(𝕀3×3𝐁j):𝐕j\displaystyle\operatorname{div}\mathscr{R}_{0}{\bf f}=\nabla\xi_{0}\cdot{\bf U}_{0}+\sum_{j=1}^{\ell}\nabla\xi_{j}\cdot{\bf V}_{j}+\sum_{j=1}^{\ell}\xi_{j}(\mathbb{I}_{3\times 3}-{\bf B}_{j})^{\top}:\nabla{\bf V}_{j}

such that

t1𝐟Lxp\displaystyle\|\partial_{t}\mathscr{R}_{1}{\bf f}\|_{L^{p}_{x}} BogΩ(ξ0t𝐔0)Lxp+j=1BogΩ(ξjt𝐕j)Lxp\displaystyle\lesssim\|\operatorname{Bog}_{{\Omega}}(\nabla\xi_{0}\cdot\partial_{t}{\bf U}_{0})\|_{L^{p}_{x}}+\sum_{j=1}^{\ell}\|\operatorname{Bog}_{{\Omega}}(\nabla\xi_{j}\cdot\partial_{t}{\bf V}_{j})\|_{L^{p}_{x}}
+j=1BogΩ(ξj(𝕀3×3𝐁j):t𝐕j)Lxp\displaystyle+\sum_{j=1}^{\ell}\|\operatorname{Bog}_{\Omega}\big{(}\xi_{j}(\mathbb{I}_{3\times 3}-\mathbf{B}_{j})^{\top}:\nabla\partial_{t}{\bf V}_{j}\big{)}\|_{L^{p}_{x}}
ξ0t𝐔0Wx1,p+j=1ξjt𝐕jWx1,p\displaystyle\lesssim\|\nabla\xi_{0}\cdot\partial_{t}{\bf U}_{0}\|_{W^{-1,p}_{x}}+\sum_{j=1}^{\ell}\|\nabla\xi_{j}\cdot\partial_{t}{\bf V}_{j}\|_{W^{-1,p}_{x}}
+j=1ξj(𝕀3×3𝐁j):t𝐕jWx1,p\displaystyle+\sum_{j=1}^{\ell}\|\xi_{j}(\mathbb{I}_{3\times 3}-{\bf B}_{j})^{\top}:\nabla\partial_{t}{\bf V}_{j}\|_{W^{-1,p}_{x}}
=:(R)1+(R)2+(R)3\displaystyle=:(R)_{1}+(R)_{2}+(R)_{3}

using continuity of the Bogovskii-operator, cf. (25). Since 𝐔0{\bf U}_{0} solves (28) we infer from (30) that

(R)1rdtt𝐔0Wx1,prdtTr/2𝐟Lxprdt.\displaystyle\int_{\mathcal{I}}(R)_{1}^{r}\,\mathrm{d}t\lesssim\int_{\mathcal{I}}\|\partial_{t}{\bf U}_{0}\|_{W^{-1,p}_{x}}^{r}\,\mathrm{d}t\lesssim T^{r/2}\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t.

Similarly, we obtain from (31)

(R)2rdtj=1t𝐕jWx1,prdtj=1t𝐔jWx1,prdtTr/2𝐟Lxprdt\displaystyle\int_{\mathcal{I}}(R)_{2}^{r}\,\mathrm{d}t\lesssim\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\partial_{t}{\bf V}_{j}\|_{W^{-1,p}_{x}}^{r}\,\mathrm{d}t\lesssim\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\partial_{t}{\bf U}_{j}\|_{W^{-1,p}_{x}}^{r}\,\mathrm{d}t\lesssim T^{r/2}\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t

using also 𝐕j=𝐔j𝚽j𝒱j{\bf V}_{j}={\bf U}_{j}\circ{\boldsymbol{\Phi}}_{j}\circ\mathscr{V}_{j} and (15) as well as (7)–(9). Finally, arguing again as in the estimates for 𝒮7\mathscr{S}_{7} above, and using div𝐁j=0\operatorname{div}\mathbf{B}_{j}=0 (which holds as a consequence of the Piola-identity)

(R)3rdt\displaystyle\int_{\mathcal{I}}(R)_{3}^{r}\,\mathrm{d}t j=1ξj(𝕀3×3𝐁j):t𝐕jWx1,prdt\displaystyle\lesssim\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\xi_{j}(\mathbb{I}_{3\times 3}-\mathbf{B}_{j})^{\top}:\nabla\partial_{t}{\bf V}_{j}\|_{W^{-1,p}_{x}}^{r}\,\mathrm{d}t
j=1𝕀3×3𝐁j1,p(Ω)rt𝐕jLxprdt\displaystyle\lesssim\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\mathbb{I}_{3\times 3}-\mathbf{B}_{j}\|_{\mathcal{M}^{1,p}({\Omega})}^{r}\|\partial_{t}{\bf V}_{j}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t
δ(Lip(Ω))𝐟Lxprdt.\displaystyle\lesssim\delta(\mathrm{Lip}(\partial{\Omega}))\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t.

In conclusion, we have shown

t1𝐟Lxprdt\displaystyle\int_{\mathcal{I}}\|\partial_{t}\mathscr{R}_{1}{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t 14𝐟Lxprdt,\displaystyle\leq\,\tfrac{1}{4}\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t, (40)

for TT and Lip(Ω)\mathrm{Lip}(\partial\Omega) sufficiently small. As far as Δ1𝐟\Delta\mathscr{R}_{1}{\bf f} is concerned, we have analogously

Δ1𝐟Lxprdt\displaystyle\int_{\mathcal{I}}\|\Delta\mathscr{R}_{1}{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t BogΩ(ξ0𝐔0)Wx2,prdt+j=1BogΩ(ξj𝐕j)Wx2,prdt\displaystyle\lesssim\int_{\mathcal{I}}\|\operatorname{Bog}_{{\Omega}}(\nabla\xi_{0}\cdot{\bf U}_{0})\|_{W^{2,p}_{x}}^{r}\,\mathrm{d}t+\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\operatorname{Bog}_{{\Omega}}(\nabla\xi_{j}\cdot{\bf V}_{j})\|_{W^{2,p}_{x}}^{r}\,\mathrm{d}t
+j=1BogΩ(ξj(𝕀3×3𝐁j):𝐕j)Wx2,prdt\displaystyle+\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\operatorname{Bog}_{\Omega}\big{(}\xi_{j}(\mathbb{I}_{3\times 3}-\mathbf{B}_{j})^{\top}:\nabla{\bf V}_{j}\big{)}\|_{W^{2,p}_{x}}^{r}\,\mathrm{d}t
ξ0𝐔0Wx1,prdt+j=1ηj𝐕jWx1,prdt\displaystyle\lesssim\int_{\mathcal{I}}\|\nabla\xi_{0}\cdot{\bf U}_{0}\|_{W^{1,p}_{x}}^{r}\,\mathrm{d}t+\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\nabla\eta_{j}\cdot{\bf V}_{j}\|_{W^{1,p}_{x}}^{r}\,\mathrm{d}t
+j=1ξj(𝕀3×3𝐁j):𝐕jWx1,prdt\displaystyle+\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\xi_{j}(\mathbb{I}_{3\times 3}-\mathbf{B}_{j})^{\top}:\nabla{\bf V}_{j}\|_{W^{1,p}_{x}}^{r}\,\mathrm{d}t
𝐔0Wx1,prdt+j=1𝐔jWx1,prdt\displaystyle\lesssim\int_{\mathcal{I}}\|{\bf U}_{0}\|_{W^{1,p}_{x}}^{r}\,\mathrm{d}t+\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|{\bf U}_{j}\|_{W^{1,p}_{x}}^{r}\,\mathrm{d}t
+j=1𝕀3×3𝐁j1,p(Ω)r𝐔jWx2,prdt\displaystyle+\sum_{j=1}^{\ell}\int_{\mathcal{I}}\|\mathbb{I}_{3\times 3}-\mathbf{B}_{j}\|^{r}_{\mathcal{M}^{1,p}({\Omega})}\|{\bf U}_{j}\|_{W^{2,p}_{x}}^{r}\,\mathrm{d}t
Tr/2𝐟Lxprdt+δ(Lip(Ω))𝐟Lxprdt.\displaystyle\lesssim T^{r/2}\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t+\delta(\mathrm{Lip}(\partial{\Omega}))\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t.

Note that we also made use of (36). This implies

Δ1𝐟Lxprdt\displaystyle\int_{\mathcal{I}}\|\Delta\mathscr{R}_{1}{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t 14𝐟Lxprdt\displaystyle\leq\,\tfrac{1}{4}\int_{\mathcal{I}}\|{\bf f}\|_{L^{p}_{x}}^{r}\,\mathrm{d}t (41)

choosing TT and Lip(Ω)\mathrm{Lip}(\partial\Omega) small enough. Combining (39), (40) and (41) implies 𝒯34\|\mathscr{T}\|\leq\tfrac{3}{4}. Hence \mathscr{L} is onto recalling (26). This means we have shown the claim for TT sufficiently small, say T=T01T=T_{0}\ll 1. It is easy to extend it to the whole interval. Let (𝐮,π)({\bf u},\pi) be the solution in [0,T][0,T]. In order to obtain a solution on the whole interval we consider a partition of unity (ψk)k=1K(\psi_{k})_{k=1}^{K} on [0,T][0,T] such that spt(ψk)(αk,αk+T0]\mathrm{spt}(\psi_{k})\subset(\alpha_{k},\alpha_{k}+T_{0}] for some (αk)k=2K[0,T](\alpha_{k})_{k=2}^{K}\subset[0,T] and α1=0\alpha_{1}=0. The functions (𝐮k,πk)=(ψk𝐮,ψkπ)({\bf u}_{k},\pi_{k})=(\psi_{k}{\bf u},\psi_{k}\pi) are the unique solutions to

t𝐮k=Δ𝐮kπk+𝐟+ψk𝐮,div𝐮k=0,𝐮k|(αk,αk+T0)×Ω=0,𝐮k(αk,)=0.\displaystyle\partial_{t}{\bf u}_{k}=\Delta{\bf u}_{k}-\nabla\pi_{k}+{\bf f}+\psi_{k}^{\prime}{\bf u},\quad\operatorname{div}{\bf u}_{k}=0,\quad{\bf u}_{k}|_{(\alpha_{k},\alpha_{k}+T_{0})\times\partial{\Omega}}=0,\quad{\bf u}_{k}(\alpha_{k},\cdot)=0.

Applying the result proved for the interval [0,T0][0,T_{0}] and noticing that 𝐮=k=1K𝐮k{\bf u}=\sum_{k=1}^{K}{\bf u}_{k} and π=k=1Kπk\pi=\sum_{k=1}^{K}\pi_{k} proves the claim in the general case. ∎

4. The perturbed system

In this section we develop a theory for some perturbed Stokes and Navier–Stokes systems which arise from the original one by flattening the boundary (introducing local coordinates as in Section 2.4). The perturbed Navier–Stokes system will be the basis for the partial regularity proof in Section 5 in which we compare its solution locally to a solution to the perturbed Stokes system. By means of Sobolev multipliers we now derive optimal assumptions concerning the coefficients in the latter allowing for a maximal regularity theory.

4.1. Perturbed Navier–Stokes equations

For a boundary suitable weak solution (𝐮,π,φ)({\bf u},\pi,\varphi) to (1) we define π¯=π𝚽\overline{\pi}=\pi\circ{\boldsymbol{\Phi}}, 𝐮¯=𝐮𝚽\overline{{\bf u}}={\bf u}\circ{\boldsymbol{\Phi}} and 𝐟¯=𝐟𝚽\overline{{\bf f}}={\bf f}\circ{\boldsymbol{\Phi}}, where 𝚽{\boldsymbol{\Phi}} is the extension of φ\varphi given in (14). We also introduce

𝐀φ\displaystyle{\bf A}_{\varphi} =Jφ(𝚿𝚽)𝚿𝚽,𝐁φ=Jφ𝚿𝚽,\displaystyle=J_{\varphi}\big{(}\nabla{\boldsymbol{\Psi}}\circ{\boldsymbol{\Phi}}\big{)}^{\top}\nabla{\boldsymbol{\Psi}}\circ{\boldsymbol{\Phi}},\quad{\bf B}_{\varphi}=J_{\varphi}\nabla{\boldsymbol{\Psi}}\circ{\boldsymbol{\Phi}}, (42)

where Jφ=det(𝚽)J_{\varphi}=\mathrm{det}(\nabla{\boldsymbol{\Phi}}). We see that (𝐮¯,π¯,φ)(\overline{\bf u},\overline{\pi},\varphi) is a solution to the system

Jφt𝐮¯+(𝐁φ𝐮¯)𝐮¯+div(𝐁φπ¯)div(𝐀φ𝐮¯)\displaystyle J_{\varphi}\partial_{t}\overline{\bf u}+({\bf B}_{\varphi}\nabla\overline{\bf u})\overline{\bf u}+\operatorname{div}\big{(}{\bf B}_{\varphi}\overline{\pi}\big{)}-\operatorname{div}\big{(}{\bf A}_{\varphi}\nabla\overline{\bf u}\big{)} =Jφ𝐟¯,\displaystyle=J_{\varphi}\overline{\bf f}, (43)
𝐁φ:𝐮¯=0,𝐮¯|1+\displaystyle{\bf B}_{\varphi}^{\top}:\nabla\overline{\bf u}=0,\quad\overline{\bf u}|_{\partial\mathcal{B}_{1}^{+}\cap\partial\mathbb{H}} =0,\displaystyle=0, (44)

a.a. in 𝒬1+\mathcal{Q}_{1}^{+}. Note that it may be necessary to translate and scale the coordinates in space-time to arrive at a system posed in 𝒬1+:=𝒬1(0,0)\mathcal{Q}_{1}^{+}:=\mathcal{Q}_{1}(0,0) (rather than in 𝒬r(t0,x0)\mathcal{Q}_{r}(t_{0},x_{0}) for some r>0r>0, t0t_{0}\in\mathcal{I} and x0x_{0}\in\partial\mathbb{H}). Similarly, we can transform the local energy inequality leading to

Ω12Jφζ|𝐮¯(t)|2dx+0tΩζ𝐀φ𝐮¯:𝐮¯dxdσ0tΩ12Jφ|𝐮¯|2tζdxdσ+0tΩ12Jφ|𝐮¯|2Δ𝚿𝚽ζdxdσ+0tΩ12|𝐮¯|2𝐀φ:2ζdxdσ+0tΩ12(|𝐮¯|2+2π¯)𝐮¯𝐁φζdxdσ+0tΩJφζ𝐟¯𝐮¯dxdσ\begin{split}\int_{\Omega}\frac{1}{2}J_{\varphi}\zeta\big{|}\overline{\bf u}(t)\big{|}^{2}\,\mathrm{d}x&+\int_{0}^{t}\int_{\Omega}\zeta{\bf A}_{\varphi}\nabla\overline{\bf u}:\nabla\overline{\bf u}\,\mathrm{d}x\,\mathrm{d}\sigma\\ &\leq\int_{0}^{t}\int_{\Omega}\frac{1}{2}J_{\varphi}|\overline{\bf u}|^{2}\partial_{t}\zeta\,\mathrm{d}x\,\mathrm{d}\sigma+\int_{0}^{t}\int_{\Omega}\frac{1}{2}J_{\varphi}|\overline{\bf u}|^{2}\Delta{\boldsymbol{\Psi}}\circ{\boldsymbol{\Phi}}\cdot\nabla\zeta\,\mathrm{d}x\,\mathrm{d}\sigma\\ &+\int_{0}^{t}\int_{\Omega}\frac{1}{2}|\overline{\bf u}|^{2}{\bf A}_{\varphi}:\nabla^{2}\zeta\,\mathrm{d}x\,\mathrm{d}\sigma+\int_{0}^{t}\int_{\Omega}\frac{1}{2}\big{(}|\overline{\bf u}|^{2}+2\overline{\pi}\big{)}\overline{\bf u}\cdot{\bf B}_{\varphi}\nabla\zeta\,\mathrm{d}x\,\mathrm{d}\sigma\\ &+\int_{0}^{t}\int_{\Omega}J_{\varphi}\zeta\overline{\bf f}\cdot\overline{\bf u}\,\mathrm{d}x\,\mathrm{d}\sigma\end{split} (45)

for any ζCc(𝒬1)\zeta\in C^{\infty}_{c}(\mathcal{Q}_{1}) with ζ0\zeta\geq 0.

Definition 4.1 (Boundary suitable weak solution perturbed system).

Let (𝐟¯,𝐮¯0)(\overline{\bf f},\overline{\bf u}_{0}) be a dataset such that

𝐟¯Lr(1;Ls(1+))),𝐮¯0W2,sW1,20,div(1+).\displaystyle\overline{\bf f}\in L^{r_{\ast}}\big{(}\mathcal{I}_{1};L^{s_{\ast}}(\mathcal{B}_{1}^{+}))\big{)},\quad\overline{\bf u}_{0}\in W^{2,s_{\ast}}\cap W^{1,2}_{0,\mathrm{\operatorname{div}}}(\mathcal{B}_{1}^{+}). (46)

We call the triple (𝐮¯,π¯,φ)(\overline{\bf u},\overline{\pi},\varphi) a boundary suitable weak solution to the perturbed Navier–Stokes system (43) with data (𝐟¯,𝐮¯0)(\overline{\bf f},\overline{\bf u}_{0}) provided that the following holds:

  • (a)

    The velocity field 𝐮¯\overline{\bf u} satisfies

    𝐮¯L(1;L2(1+))L2(1;Wdiv1,2(1+))Lr(1;W2,s(1+))W1,r(1;Ls(1+)).\displaystyle\overline{\bf u}\in L^{\infty}\big{(}\mathcal{I}_{1};L^{2}(\mathcal{B}_{1}^{+})\big{)}\cap L^{2}\big{(}\mathcal{I}_{1};W^{1,2}_{\operatorname{div}}(\mathcal{B}_{1}^{+})\big{)}\cap L^{r_{\ast}}(\mathcal{I}_{1};W^{2,s_{\ast}}(\mathcal{B}_{1}^{+}))\cap W^{1,r_{\ast}}(\mathcal{I}_{1};L^{s_{\ast}}(\mathcal{B}_{1}^{+})).
  • (b)

    The pressure π¯\overline{\pi} satisfies

    π¯Lr(1;W1,s(1+)).\overline{\pi}\in L^{r_{\ast}}(\mathcal{I}_{1};W^{1,s_{\ast}}_{\perp}(\mathcal{B}_{1}^{+})).
  • c)

    The boundary coordinates satisfy (19).

  • (d)

    The system (43)–(44) holds a.a. in 𝒬1+\mathcal{Q}^{+}_{1}.

  • (e)

    The local energy inequality (45) holds for any ζCc(𝒬1)\zeta\in C_{c}^{\infty}(\mathcal{Q}_{1}) with ζ0\zeta\geq 0.

A crucial part of the partial regularity proof in Section 5 will be the comparison of a boundary suitable weak solution of the perturbed Navier–Stokes system with a solution of the perturbed Stokes system. The analysis of the latter is the content of the following subsection.

4.2. Perturbed Stokes equations

In this section we consider, in analogy to (43)–(44), a perturbed Stokes system in 𝒬1+\mathcal{Q}_{1}^{+} of the form

Jφt𝐮¯+div(𝐁φπ¯)div(𝐀φ𝐮¯)=𝐠¯,𝐁φ:𝐮¯=h¯,𝐮¯|1+=0,𝐮¯(1,)=0,\displaystyle\begin{aligned} J_{\varphi}\partial_{t}\overline{\bf u}+\operatorname{div}\big{(}{\bf B}_{\varphi}\overline{\pi}\big{)}-\operatorname{div}\big{(}{\bf A}_{\varphi}\nabla\overline{\bf u}\big{)}&=\overline{\bf g},\\ {\bf B}_{\varphi}^{\top}:\nabla\overline{{\bf u}}=\overline{h},\quad\overline{\bf u}|_{\mathcal{B}^{+}_{1}\cap\partial\mathbb{H}}&=0,\quad\overline{{\bf u}}(-1,\cdot)=0,\end{aligned} (47)

where 𝐀φ{\bf A}_{\varphi} and 𝐁φ{\bf B}_{\varphi} are given in accordance with (42) for a given function φ:2\varphi:\mathbb{R}^{2}\rightarrow\mathbb{R} and 𝐠¯\overline{\bf g} and h¯\overline{h} are given data. We obtain the following maximal regularity result.

Lemma 4.2.

Let p,r(1,)p,r\in(1,\infty). Suppose that φ21/p,p(2)(δ)\varphi\in\mathcal{M}^{2-1/p,p}(\mathbb{R}^{2})(\delta) and that φWy1,δ\|\varphi\|_{W^{1,\infty}_{y}}\leq\delta for some sufficiently small δ\delta. Assume further that 𝐠¯Lr(1;Lp(1+))\overline{\bf g}\in L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1}^{+})) and h¯Lr(1;W1,pLp(1+))\overline{h}\in L^{r}(\mathcal{I}_{1};W^{1,p}\cap L^{p}_{\perp}(\mathcal{B}_{1}^{+})) with th¯Lr(1;W1,p(1+))\partial_{t}\overline{h}\in L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}_{1}^{+})) and h¯(0,)=0\overline{h}(0,\cdot)=0. Then there is a unique solution (𝐮¯,π¯)(\overline{\bf u},\overline{\pi}) to (47) which satisfies

t𝐮¯Lr(1;Lp(1+))+𝐮¯Lr(1;W2,p(1+))+π¯Lr(1;W1,p(1+))𝐠¯Lr(1;Lp(1+))+h¯Lr(1;Lp(1+))+th¯Lr(1;W1,p(1+)),\displaystyle\begin{aligned} \|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}&+\|\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};W^{2,p}(\mathcal{B}^{+}_{1}))}+\|\overline{\pi}\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}^{+}_{1}))}\\ &\lesssim\|\overline{\bf g}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\nabla\overline{h}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\partial_{t}\overline{h}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))},\end{aligned} (48)

where the hidden constant only depends on p,rp,r and δ\delta.

Proof.

Let us initially assume that 𝐮¯\overline{\bf u} and π¯\overline{\pi} are sufficiently smooth. We rewrite (47) as

t𝐮¯+π¯Δ𝐮¯=𝐠¯+(1Jφ)t𝐮¯+div((𝕀3×3𝐁φ)π¯)+div((𝐀φ𝕀3×3)𝐮¯),div𝐮¯=(𝕀3×3𝐁φ):𝐮¯+h¯,𝐮¯|1+=0,𝐮¯(1,)=0.\displaystyle\begin{aligned} \partial_{t}\overline{\bf u}+\nabla\overline{\pi}-\Delta\overline{\bf u}&=\overline{\bf g}+(1-J_{\varphi})\partial_{t}\overline{\bf u}\\ &+\operatorname{div}\big{(}(\mathbb{I}_{3\times 3}-{\bf B}_{\varphi})\overline{\pi}\big{)}+\operatorname{div}\big{(}({\bf A}_{\varphi}-\mathbb{I}_{3\times 3})\nabla\overline{\bf u}\big{)},\\ \operatorname{div}\overline{{\bf u}}=\big{(}\mathbb{I}_{3\times 3}-{\bf B}_{\varphi}\big{)}^{\top}&:\nabla\overline{{\bf u}}+\overline{h},\quad\overline{\bf u}|_{\mathcal{B}^{+}_{1}\cap\partial\mathbb{H}}=0,\quad\overline{{\bf u}}(-1,\cdot)=0.\end{aligned} (49)

This is a classical Stokes system in 𝒬1+\mathcal{Q}_{1}^{+} with right-hand side

𝐆¯:=𝐠¯+(1Jφ)t𝐮¯+div((𝕀3×3𝐁φ)π¯)+div((𝐀φ𝕀3×3)𝐮¯)\displaystyle\overline{\bf G}:=\overline{\bf g}+(1-J_{\varphi})\partial_{t}\overline{\bf u}+\operatorname{div}\big{(}(\mathbb{I}_{3\times 3}-{\bf B}_{\varphi})\overline{\pi}\big{)}+\operatorname{div}\big{(}({\bf A}_{\varphi}-\mathbb{I}_{3\times 3})\nabla\overline{\bf u}\big{)}

and given divergence

H¯:=(𝕀3×3𝐁φ)\displaystyle\overline{H}:=\big{(}\mathbb{I}_{3\times 3}-{\bf B}_{\varphi}\big{)}^{\top} :𝐮¯+h¯.\displaystyle:\nabla\overline{{\bf u}}+\overline{h}.

The known regularity theory (see [12] or [32]) yields

t𝐮¯Lr(1;Lp(1+))+𝐮¯Lr(1;W2,p(1+))+π¯Lr(1;W1,p(1+))𝐆¯Lr(1;Lp(1+))+H¯Lr(1;Lp(1+))+tH¯Lr(1;W1,p(1+)).\displaystyle\begin{aligned} \|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}&+\|\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};W^{2,p}(\mathcal{B}^{+}_{1}))}+\|\overline{\pi}\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}^{+}_{1}))}\\ &\lesssim\|\overline{\bf G}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\nabla\overline{H}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\partial_{t}\overline{H}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))}.\end{aligned} (50)

The goal is now to estimate the norms of 𝐆¯\overline{\bf G} and H¯\overline{H} employing the theory of Sobolev multipliers. First of all, we deduce from (12) and (15) that

(1Jφ)t𝐮¯Lr(1;Lp(1+)δt𝐮¯Lr(1;Lp(1+))\displaystyle\|(1-J_{\varphi})\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1})}\lesssim\delta\|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}

using the assumption of a small Lipschitz constant.

Arguing similarly to the proof of Theorem 3.1 we have

div((𝐀φ𝕀3×3)𝐮¯)Lp(1+)\displaystyle\|\operatorname{div}\big{(}({\bf A}_{\varphi}-\mathbb{I}_{3\times 3})\nabla\overline{\bf u}\big{)}\|_{L^{p}(\mathcal{B}_{1}^{+})} 𝒯φ(W2,p(1+))𝐮¯W2,p(1+)\displaystyle\lesssim\|\mathcal{T}\varphi\|_{\mathcal{M}(W^{2,p}(\mathcal{B}_{1}^{+}))}\|\overline{{\bf u}}\|_{W^{2,p}(\mathcal{B}_{1}^{+})}

as well as

div((𝕀3×3𝐁φ)π¯)Lp(1+)\displaystyle\|\operatorname{div}\big{(}(\mathbb{I}_{3\times 3}-{\bf B}_{\varphi})\overline{\pi}\big{)}\|_{L^{p}(\mathcal{B}_{1}^{+})} 𝐁φ𝕀3×31,p(1+)π¯W1,p(1+)\displaystyle\lesssim\|{\bf B}_{\varphi}-\mathbb{I}_{3\times 3}\|_{\mathcal{M}^{1,p}(\mathcal{B}_{1}^{+})}\|\overline{\pi}\|_{W^{1,p}(\mathcal{B}_{1}^{+})}
𝒯φ2,p()π¯W1,p(1+).\displaystyle\lesssim\|\mathcal{T}\varphi\|_{\mathcal{M}^{2,p}(\mathbb{H})}\|\overline{\pi}\|_{W^{1,p}(\mathcal{B}_{1}^{+})}.

By (13) we have

𝒯φ2,p()φ21/p,p(n1)δ\displaystyle\|\mathcal{T}\varphi\|_{\mathcal{M}^{2,p}(\mathbb{H})}\lesssim\|\varphi\|_{\mathcal{M}^{2-1/p,p}(\mathbb{R}^{n-1})}\leq\delta (51)

using our assumption in the last step. We conclude that

𝐆Lr(1;Lp(1+))\displaystyle\|{\bf G}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))} 𝐠¯Lr(1;Lp(1+))+δ(t𝐮¯Lr(1;Lp(1+))+𝐮¯Lr(1;W2,p(1+)))\displaystyle\lesssim\|\overline{\bf g}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\delta\big{(}\|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};W^{2,p}(\mathcal{B}^{+}_{1}))}\big{)}
+δπ¯Lr(1;W1,p(1+)).\displaystyle+\delta\|\overline{\pi}\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}^{+}_{1}))}.

By analogous arguments we obtain

((𝕀3×3𝐁φ):𝐮¯)Lp(1+)\displaystyle\|\nabla\big{(}(\mathbb{I}_{3\times 3}-{\bf B}_{\varphi})^{\top}:\nabla\overline{{\bf u}}\big{)}\|_{L^{p}(\mathcal{B}_{1}^{+})} 𝐁φ𝕀3×31,p(1+)𝐮¯W1,p(1+)\displaystyle\lesssim\|{\bf B}_{\varphi}-\mathbb{I}_{3\times 3}\|_{\mathcal{M}^{1,p}(\mathcal{B}_{1}^{+})}\|\nabla\overline{\bf u}\|_{W^{1,p}(\mathcal{B}_{1}^{+})}
𝒯φ2,p()𝐮¯W2,p(1+),\displaystyle\lesssim\|\mathcal{T}\varphi\|_{\mathcal{M}^{2,p}(\mathbb{H})}\|\overline{\bf u}\|_{W^{2,p}(\mathcal{B}_{1}^{+})},
t((𝕀3×3𝐁φ):𝐮¯)W1,p(1+)\displaystyle\|\partial_{t}\big{(}(\mathbb{I}_{3\times 3}-{\bf B}_{\varphi})^{\top}:\nabla\overline{{\bf u}}\big{)}\|_{W^{-1,p}(\mathcal{B}_{1}^{+})} =(𝕀3×3𝐁φ):t𝐮¯W1,p(1+)\displaystyle=\|(\mathbb{I}_{3\times 3}-{\bf B}_{\varphi})^{\top}:\nabla\partial_{t}\overline{{\bf u}}\|_{W^{-1,p}(\mathcal{B}_{1}^{+})}
𝐁φ𝕀3×31,p(1+)t𝐮¯W1,p(1+)\displaystyle\leq\|{\bf B}_{\varphi}-\mathbb{I}_{3\times 3}\|_{\mathcal{M}^{1,p}(\mathcal{B}_{1}^{+})}\|\partial_{t}\nabla\overline{\bf u}\|_{W^{-1,p}(\mathcal{B}_{1}^{+})}
𝒯φ2,p()t𝐮¯Lp(1+),\displaystyle\lesssim\|\mathcal{T}\varphi\|_{\mathcal{M}^{2,p}(\mathbb{H})}\|\partial_{t}\overline{\bf u}\|_{L^{p}(\mathcal{B}_{1}^{+})},

such that

H¯Lr(1;Lp(1+))+tH¯Lr(1;W1,p(1+))\displaystyle\|\nabla\overline{H}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\partial_{t}\overline{H}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))} h¯Lr(1;Lp(1+))+th¯Lr(1;W1,p(1+))\displaystyle\lesssim\|\nabla\overline{h}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\partial_{t}\overline{h}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))}
+δ(t𝐮¯Lr(1;Lp(1+))+𝐮¯Lr(1;W2,p(1+))).\displaystyle+\delta\big{(}\|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};W^{2,p}(\mathcal{B}^{+}_{1}))}\big{)}.

Combing the estimates for 𝐆¯\overline{\bf G} and H¯\overline{H} yields the claim for δ\delta sufficiently small as a consequence of (50).

Since smoothness of (𝐮¯,π¯)(\overline{\bf u},\overline{\pi}) is not a priori known one has to regularise the equations. This can be done by mollifying the function φ\varphi with a standard mollifier. As shown in [19, Lemma 4.3.3.] mollification does not expand the s,p(2)\mathcal{M}^{s,p}(\mathbb{R}^{2})-norm. For the regularised problem the results from [26, Lemma 3.1] apply and we obtain a sufficiently smooth solution. The previous estimates can then be performed uniformly with respect to the mollification parameter and the claimed result follows in the limit. ∎

As in [26, Lemma 3.2] we deduce the following Cacciopoli-type inequality from Lemma 4.2.

Lemma 4.3.

Let p,q,r(1,)p,q,r\in(1,\infty) with qpq\geq p. Suppose that φ21/q,q(2)(δ)\varphi\in\mathcal{M}^{2-1/q,q}(\mathbb{R}^{2})(\delta) and that φWy1,δ\|\varphi\|_{W^{1,\infty}_{y}}\leq\delta for some sufficiently small δ\delta. Assume further that 𝐠¯Lr(1;Lp(1+))\overline{\bf g}\in L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1}^{+})) and that h¯=0\overline{h}=0. The solution (𝐮¯,π¯)(\overline{\bf u},\overline{\pi}) to (47) satisfies

t𝐮¯Lr(1/2;Lq(1/2+))+𝐮¯Lr(1/2;W2,q(1/2+))+π¯Lr(1/2;W1,q(1/2+))𝐠¯Lr(1;Lq(1+))+𝐮¯Lr(1;Lp(1+))+π¯(π¯)1+Lr(1;Lp(1+)),\displaystyle\begin{aligned} \|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1/2};L^{q}(\mathcal{B}^{+}_{1/2}))}&+\|\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1/2};W^{2,q}(\mathcal{B}^{+}_{1/2}))}+\|\overline{\pi}\|_{L^{r}(\mathcal{I}_{1/2};W^{1,q}(\mathcal{B}^{+}_{1/2}))}\\ &\lesssim\|\overline{\bf g}\|_{L^{r}(\mathcal{I}_{1};L^{q}(\mathcal{B}^{+}_{1}))}+\|\nabla\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\overline{\pi}-(\overline{\pi})_{\mathcal{B}_{1}^{+}}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))},\end{aligned} (52)

where the hidden constant only depends on p,q,rp,q,r and δ\delta.

Proof.

Let us initially suppose that p=qp=q. We consider a cut-off function ζCc(𝒬1)\zeta\in C^{\infty}_{c}(\mathcal{Q}_{1}) with ζ=1\zeta=1 in 𝒬3/4\mathcal{Q}_{3/4}. The functions 𝐯¯=ζ𝐮¯\overline{\bf v}=\zeta\overline{\bf u} and 𝔮¯=ζπ¯\overline{\mathfrak{q}}=\zeta\overline{\pi} solve the system

Jφt𝐯¯+div(𝐁φπ¯)div(𝐀φ𝐯¯)=𝐆¯ζ,𝐁φ:𝐯¯=h¯ζ,𝐯¯|1+=0,𝐯¯(1,)=0,\displaystyle\begin{aligned} J_{\varphi}\partial_{t}\overline{\bf v}+\operatorname{div}\big{(}{\bf B}_{\varphi}\overline{\pi}\big{)}-\operatorname{div}\big{(}{\bf A}_{\varphi}\nabla\overline{\bf v}\big{)}&=\overline{\bf G}_{\zeta},\\ {\bf B}_{\varphi}:\nabla\overline{{\bf v}}=\overline{h}_{\zeta},\quad\overline{\bf v}|_{\mathcal{B}^{+}_{1}\cap\partial\mathbb{H}}&=0,\quad\overline{{\bf v}}(-1,\cdot)=0,\end{aligned} (53)

where 𝐆¯ζ\overline{\bf G}_{\zeta} and h¯ζ\overline{h}_{\zeta} are given by

𝐆¯ζ\displaystyle\overline{\bf G}_{\zeta} =ζ𝐠¯Jφtζ𝐮¯2𝐀φ𝐮¯ζ+(π¯(π¯)1+)𝐁φζ𝐮¯div(𝐀φζ),h¯ζ=𝐮¯𝐁φζ.\displaystyle=\zeta\overline{\bf g}-J_{\varphi}\partial_{t}\zeta\overline{\bf u}-2{\bf A}_{\varphi}\nabla\overline{\bf u}\nabla\zeta+(\overline{\pi}-(\overline{\pi})_{\mathcal{B}_{1}^{+}}){\bf B}_{\varphi}\nabla\zeta-\overline{\bf u}\operatorname{div}\big{(}{\bf A}_{\varphi}\nabla\zeta),\quad\overline{h}_{\zeta}=\overline{\bf u}\cdot{\bf B}_{\varphi}\nabla\zeta.

We apply Lemma 4.2 to (53) and obtain

t𝐯¯Lr(1;Lp(1+))\displaystyle\|\partial_{t}\overline{\bf v}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))} +𝐯¯Lr(1;W2,p(1+))+𝔮¯Lr(1;W1,p(1+))\displaystyle+\|\overline{\bf v}\|_{L^{r}(\mathcal{I}_{1};W^{2,p}(\mathcal{B}^{+}_{1}))}+\|\overline{\mathfrak{q}}\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}^{+}_{1}))}
𝐆¯ζLr(1;Lp(1+))+h¯ζLr(1;Lp(1+))+th¯ζLr(1;W1,p(1+)).\displaystyle\lesssim\|\overline{\bf G}_{\zeta}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\nabla\overline{h}_{\zeta}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\partial_{t}\overline{h}_{\zeta}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))}.

We clearly have

𝐆¯ζLr(1;Lp(1+))\displaystyle\|\overline{\bf G}_{\zeta}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))} 𝐠¯Lr(1;Lp(1+))+𝐮¯Lr(1;Lp(1+))+π¯(π¯)1+Lr(1;Lp(1+))\displaystyle\lesssim\|\overline{\bf g}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\nabla\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\overline{\pi}-(\overline{\pi})_{\mathcal{B}_{1}^{+}}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}
+𝐮¯div(𝐀φζ)Lr(1;Lp(1+))\displaystyle+\|\overline{\bf u}\operatorname{div}\big{(}{\bf A}_{\varphi}\nabla\zeta)\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}

since φ\varphi is Lipschitz by assumption (and so are 𝚽{\boldsymbol{\Phi}} and 𝚿{\boldsymbol{\Psi}}, cf. (12) and (16)). Note that we also used Poincaré’s inequality recalling that 𝐮¯|1+=0\overline{\bf u}|_{\mathcal{B}^{+}_{1}\cap\partial\mathbb{H}}=0. For the last term in the above we have

𝐮¯div(𝐀φζ)Lr(1;Lp(1+))\displaystyle\|\overline{\bf u}\operatorname{div}\big{(}{\bf A}_{\varphi}\nabla\zeta)\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))} div(𝐮¯𝐀φζ)Lr(1;Lp(1+))+𝐮¯𝐀φζLr(1;Lp(1+))\displaystyle\lesssim\|\operatorname{div}\big{(}\overline{\bf u}\otimes{\bf A}_{\varphi}\nabla\zeta)\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\nabla\overline{\bf u}\,{\bf A}_{\varphi}\nabla\zeta\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}
𝐮¯𝐀φζLr(1;W1,p(1+))+𝐮¯Lr(1;Lp(1+))\displaystyle\lesssim\|\overline{\bf u}\otimes{\bf A}_{\varphi}\nabla\zeta\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}^{+}_{1}))}+\|\nabla\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}
𝐀φ1,p(1+)𝐮¯ζLr(1;W1,p(1+))+𝐮¯Lr(1;Lp(1+))\displaystyle\lesssim\|{\bf A}_{\varphi}\|_{\mathcal{M}^{1,p}(\mathcal{B}^{+}_{1})}\|\overline{\bf u}\otimes\nabla\zeta\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}^{+}_{1}))}+\|\nabla\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}
𝐮¯Lr(1;Lp(1+)).\displaystyle\lesssim\|\nabla\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}.

Note that we used in the last step, in addition to the Lipschitz continuity of 𝚽{\boldsymbol{\Phi}} and 𝚿{\boldsymbol{\Psi}}, that 𝚽{\boldsymbol{\Phi}} and 𝚿{\boldsymbol{\Psi}} belong to the correct mulitplier space by (5) and (17) using the assumption φ21/p,p(2)(δ)\varphi\in\mathcal{M}^{2-1/p,p}(\mathbb{R}^{2})(\delta). Similarly, we have

h¯Lr(1;W1,p(1+))\displaystyle\|\overline{h}\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}^{+}_{1}))} 𝐮¯𝐁φζLr(1;W1,p(1+))\displaystyle\lesssim\|\overline{\bf u}\cdot{\bf B}_{\varphi}\nabla\zeta\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}^{+}_{1}))}
𝐁ζ1,p(1+)𝐮¯ζLr(1;W1,p(1+))\displaystyle\lesssim\|{\bf B}_{\zeta}\|_{\mathcal{M}^{1,p}(\mathcal{B}^{+}_{1})}\|\overline{\bf u}\otimes\nabla\zeta\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}^{+}_{1}))}
𝐮¯Lr(1;Lp(1+)).\displaystyle\lesssim\|\nabla\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}.

Finally,

th¯Lr(1;W1,p(1+))\displaystyle\|\partial_{t}\overline{h}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))} t𝐮¯Lr(1;W1,p(1+))+𝐮¯Lr(1;W1,p(1+)).\displaystyle\lesssim\|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))}+\|\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))}.

Using equation (53) together with strict positivity of JφJ_{\varphi} and Jφ1,p(1+)J_{\varphi}\in\mathcal{M}^{1,p}(\mathcal{B}^{+}_{1}) we have

t𝐮¯Lr(1;W1,p(1+))\displaystyle\|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))} 𝐆¯Lr(1;W1,p(1+))+div(𝐁φπ¯)Lr(1;W1,p(1+))\displaystyle\lesssim\|\overline{\bf G}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))}+\|\operatorname{div}\big{(}{\bf B}_{\varphi}\overline{\pi}\big{)}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))}
+div(𝐀φ𝐮¯)Lr(1;W1,p(1+))\displaystyle+\|\operatorname{div}\big{(}{\bf A}_{\varphi}\nabla\overline{\bf u}\big{)}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}^{+}_{1}))}
𝐆¯Lr(1;Lp(1+))+𝐮¯Lr(1;Lp(1+))+π¯(π¯)1+Lr(1;Lp(1+)).\displaystyle\lesssim\|\overline{\bf G}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\nabla\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}+\|\overline{\pi}-(\overline{\pi})_{\mathcal{B}_{1}^{+}}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1}))}.

Note that we used again boundedness of 𝐀φ{\bf A}_{\varphi} and 𝐁φ{\bf B}_{\varphi}. Combining everything, we have proved (52) for the case p=qp=q (even with norms over 𝒬3/4+\mathcal{Q}^{+}_{3/4} on the left-hand side).

Let us now consider the case q(p,3p3p]q\in(p,\tfrac{3p}{3-p}] for the case p<3p<3 and q(p,)q\in(p,\infty) arbitrary for p3p\geq 3. We make the same definitions of 𝐯¯\overline{\bf v} and 𝔮¯\overline{\mathfrak{q}} as above but consider a cut-off function ζCc(𝒬3/4)\zeta\in C^{\infty}_{c}(\mathcal{Q}_{3/4}) with ζ=1\zeta=1 in 𝒬1/2\mathcal{Q}_{1/2}. By the choice of qq and Sobolev’s embedding W1,p(3/4+)Lq(3/4+)W^{1,p}(\mathcal{B}_{3/4}^{+})\hookrightarrow L^{q}(\mathcal{B}_{3/4}^{+}) we have

𝐆¯ζLr(3/4;Lq(3/4+))\displaystyle\|\overline{\bf G}_{\zeta}\|_{L^{r}(\mathcal{I}_{3/4};L^{q}(\mathcal{B}^{+}_{3/4}))} 𝐠¯Lr(3/4;Lq(3/4+))+𝐮¯Lr(3/4;Lq(3/4+))\displaystyle\lesssim\|\overline{\bf g}\|_{L^{r}(\mathcal{I}_{3/4};L^{q}(\mathcal{B}^{+}_{3/4}))}+\|\nabla\overline{\bf u}\|_{L^{r}(\mathcal{I}_{3/4};L^{q}(\mathcal{B}^{+}_{3/4}))}
+π¯(π¯)3/4+Lr(3/4;Lq(3/4+))\displaystyle+\|\overline{\pi}-(\overline{\pi})_{\mathcal{B}_{3/4}^{+}}\|_{L^{r}(\mathcal{I}_{3/4};L^{q}(\mathcal{B}^{+}_{3/4}))}
𝐠¯Lr(3/4;Lq(3/4+))+2𝐮¯Lr(3/4;Lp(3/4+))+π¯Lr(3/4;Lp(3/4+)),\displaystyle\lesssim\|\overline{\bf g}\|_{L^{r}(\mathcal{I}_{3/4};L^{q}(\mathcal{B}^{+}_{3/4}))}+\|\nabla^{2}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{3/4};L^{p}(\mathcal{B}^{+}_{3/4}))}+\|\nabla\overline{\pi}\|_{L^{r}(\mathcal{I}_{3/4};L^{p}(\mathcal{B}^{+}_{3/4}))},

arguing as for the LpL^{p}-estimates above (but using φ21/q,q(2)(δ)\varphi\in\mathcal{M}^{2-1/q,q}(\mathbb{R}^{2})(\delta)). Similarly, we obtain

h¯ζLr(3/4;Lq(3/4+))\displaystyle\|\nabla\overline{h}_{\zeta}\|_{L^{r}(\mathcal{I}_{3/4};L^{q}(\mathcal{B}^{+}_{3/4}))} 2𝐮¯Lr(3/4;Lp(3/4+)),\displaystyle\lesssim\|\nabla^{2}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{3/4};L^{p}(\mathcal{B}^{+}_{3/4}))},
th¯ζLr(3/4;W1,q(3/4+))\displaystyle\|\partial_{t}\overline{h}_{\zeta}\|_{L^{r}(\mathcal{I}_{3/4};W^{-1,q}(\mathcal{B}^{+}_{3/4}))} t𝐮¯Lr(3/4;Lp(3/4+))+𝐮¯Lr(3/4;Lp(3/4+)).\displaystyle\lesssim\|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{3/4};L^{p}(\mathcal{B}^{+}_{3/4}))}+\|\overline{\bf u}\|_{L^{r}(\mathcal{I}_{3/4};L^{p}(\mathcal{B}^{+}_{3/4}))}.

Applying now the result from the first step as well as Lemma 4.2 (but with the stronger assumption φ21/q,q(2)(δ)\varphi\in\mathcal{M}^{2-1/q,q}(\mathbb{R}^{2})(\delta), see [19, Proposition 3.5.3.] for inclusions of the spaces s,p\mathcal{M}^{s,p}) yields again the claim. Note that we still have the restriction q3p3pq\leq\tfrac{3p}{3-p} if p<3p<3. Iterating this argument based on Sobolev’s embedding and consider appropriate cut-off functions clearly allows any choice of exponent q<q<\infty. ∎

We will also need interior estimates for points close to the boundary which is why we consider in analogy to (47) the system

Jφt𝐮¯+div(𝐁φπ¯)div(𝐀φ𝐮¯)=𝐠,𝐁φ:𝐮¯=h,𝐮¯|1=0,𝐮¯(1,)=0,\displaystyle\begin{aligned} J_{\varphi}\partial_{t}\overline{\bf u}+\operatorname{div}\big{(}{\bf B}_{\varphi}\overline{\pi}\big{)}-\operatorname{div}\big{(}{\bf A}_{\varphi}\nabla\overline{\bf u}\big{)}&={\bf g},\\ {\bf B}_{\varphi}^{\top}:\nabla\overline{{\bf u}}=h,\quad\overline{\bf u}|_{\mathcal{B}_{1}}&=0,\quad\overline{{\bf u}}(-1,\cdot)=0,\end{aligned} (54)

in 𝒬1\mathcal{Q}_{1}. Arguing exactly as for Lemmas 4.2 and 4.3 we obtain the following results.

Lemma 4.4.

Let p,r(1,)p,r\in(1,\infty). Suppose that φ21/p,p(2)(δ)\varphi\in\mathcal{M}^{2-1/p,p}(\mathbb{R}^{2})(\delta) and that φWy1,δ\|\varphi\|_{W^{1,\infty}_{y}}\leq\delta for some sufficiently small δ\delta. Assume further that 𝐠¯Lr(1;Lp(1))\overline{\bf g}\in L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1})) and h¯Lr(1;W1,p(1))\overline{h}\in L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}_{1})) with th¯Lr(1;W1,p(1))\partial_{t}\overline{h}\in L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}_{1})). Then there is a unique solution (𝐮¯,π¯)(\overline{\bf u},\overline{\pi}) to (54) which satisfies

t𝐮¯Lr(1;Lp(1))+𝐮¯Lr(1;W2,p(1))+π¯Lr(1;W1,p(1))𝐠¯Lr(1;Lp(1))+h¯Lr(1;Lp(1))+th¯Lr(1;W1,p(1)),\displaystyle\begin{aligned} \|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1}))}&+\|\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};W^{2,p}(\mathcal{B}_{1}))}+\|\overline{\pi}\|_{L^{r}(\mathcal{I}_{1};W^{1,p}(\mathcal{B}_{1}))}\\ &\lesssim\|\overline{\bf g}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1}))}+\|\nabla\overline{h}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1}))}+\|\partial_{t}\overline{h}\|_{L^{r}(\mathcal{I}_{1};W^{-1,p}(\mathcal{B}_{1}))},\end{aligned} (55)

where the hidden constant only depends on p,rp,r and δ\delta.

Lemma 4.5.

Let p,q,r(1,)p,q,r\in(1,\infty) with qpq\geq p. Suppose that φ21/q,q(2)(δ)\varphi\in\mathcal{M}^{2-1/q,q}(\mathbb{R}^{2})(\delta) and that φWy1,δ\|\varphi\|_{W^{1,\infty}_{y}}\leq\delta for some sufficiently small δ\delta. Assume further that 𝐠¯Lr(1;Lp(1))\overline{\bf g}\in L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1})) and that h¯=0\overline{h}=0. The solution (𝐮¯,π¯)(\overline{\bf u},\overline{\pi}) to (54) satisfies

t𝐮¯Lr(1/2;Lq(1/2))+𝐮¯Lr(1/2;W2,q(1/2))+π¯Lr(1/2;W1,q(1/2))𝐠¯Lr(1;Lq(1))+𝐮¯Lr(1;Lp(1))+π¯(π¯)1Lr(1;Lp(1)),\displaystyle\begin{aligned} \|\partial_{t}\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1/2};L^{q}(\mathcal{B}_{1/2}))}&+\|\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1/2};W^{2,q}(\mathcal{B}_{1/2}))}+\|\overline{\pi}\|_{L^{r}(\mathcal{I}_{1/2};W^{1,q}(\mathcal{B}_{1/2}))}\\ &\lesssim\|\overline{\bf g}\|_{L^{r}(\mathcal{I}_{1};L^{q}(\mathcal{B}_{1}))}+\|\nabla\overline{\bf u}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1}))}+\|\overline{\pi}-(\overline{\pi})_{\mathcal{B}_{1}}\|_{L^{r}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1}))},\end{aligned} (56)

where the hidden constant only depends on p,q,rp,q,r and δ\delta.

5. Blow-up & proof of Theorem 2.6

We consider a boundary suitable weak solution (𝐮¯,π¯,φ)(\overline{\bf u},\overline{\pi},\varphi) to the perturbed Navier–Stokes system (43)–(44) with forcing 𝐟¯\overline{\bf f} as defined in Definition 4.1. We define the excess-functional as

rt0,x0(𝐮¯,π¯)\displaystyle\mathscr{E}_{r}^{t_{0},x_{0}}(\overline{\bf u},\overline{\pi}) :=𝒬r+(t0,x0)|𝐮¯|3dydσ+r3(𝒬r+(t0,x0)|π¯(π¯)r+(x0)|5/3dydσ)95\displaystyle:=\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{r}(t_{0},x_{0})}|\overline{\bf u}|^{3}\,\mathrm{d}y\,\mathrm{d}\sigma+r^{3}\bigg{(}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{r}(t_{0},x_{0})}|\overline{\pi}-(\overline{\pi})_{\mathcal{B}^{+}_{r}(x_{0})}|^{5/3}\,\mathrm{d}y\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}}

and use the short-hand notation r(𝐮¯,π¯)=:r0,0(𝐮¯,π¯)\mathscr{E}_{r}(\overline{\bf u},\overline{\pi})=:\mathscr{E}_{r}^{0,0}(\overline{\bf u},\overline{\pi}). The following lemma is the bulk of the partial regularity proof and the claim of Theorem 2.6 will follow in a standard manner.

Lemma 5.1.

Let p>154p>\frac{15}{4} and 𝐟¯Lp(1;Lp(1+))\overline{\bf f}\in L^{p}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1}^{+})) be given. Suppose that δ>0\delta>0 is sufficienlty small. For any τ(0,1/2)\tau\in(0,1/2) there exist constants ε>0\varepsilon>0 (small) and C>0C_{\ast}>0 (large) such the following implication is true for any triple (𝐮¯,π¯,φ)(\overline{\bf u},\overline{\pi},\varphi) which is a boundary suitable weak solution to (43)–(44) in 𝒬1+\mathcal{Q}_{1}^{+} in the sense of Definition 2.4, where φ21/p,p(2)(δ)\varphi\in\mathcal{M}^{2-1/p,p}(\mathbb{R}^{2})(\delta) with φWy1,δ\|\varphi\|_{W^{1,\infty}_{y}}\leq\delta: Suppose that

1(𝐮¯,π¯)+(𝒬1+|𝐟¯|pdydσ)3pε,\mathscr{E}_{1}(\overline{\bf u},\overline{\pi})+\bigg{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\mathcal{Q}^{+}_{1}}|\overline{\bf f}|^{p}\,\mathrm{d}y\,\mathrm{d}\sigma\bigg{)}^{\frac{3}{p}}\leq\varepsilon, (57)

then we have

τ(𝐮¯,π¯)Cτ2α(1(𝐮¯,π¯)+(𝒬1+|𝐟¯|pdydσ)3p),\mathscr{E}_{\tau}(\overline{\bf u},\overline{\pi})\leq C_{\ast}\tau^{2\alpha}\bigg{(}\mathscr{E}_{1}(\overline{\bf u},\overline{\pi})+\bigg{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\mathcal{Q}^{+}_{1}}|\overline{\bf f}|^{p}\,\mathrm{d}y\,\mathrm{d}\sigma\bigg{)}^{\frac{3}{p}}\bigg{)}, (58)

where α=3(2532p)\alpha=3\big{(}\frac{2}{5}-\frac{3}{2p}\big{)}.

Proof.

We argue by contradiction. Suppose there is τ(0,12)\tau\in(0,\tfrac{1}{2}), and a sequence of boundary suitable weak solution (𝐮¯m,π¯m,φm)(\overline{\bf u}_{m},\overline{\pi}_{m},\varphi_{m}) such that

φmLy+φm21/p,p(2)δ,φm(0)=0,φm(0)=0,\displaystyle\|\nabla\varphi_{m}\|_{L^{\infty}_{y}}+\|\varphi_{m}\|_{\mathcal{M}^{2-1/p,p}(\mathbb{R}^{2})}\leq\delta,\quad\varphi_{m}(0)=0,\quad\nabla\varphi_{m}(0)=0, (59)

as well as

λm3:=\displaystyle\lambda_{m}^{3}:= 1(𝐮¯m,π¯m)+𝐟¯mLp(1;Lp(1+))30,m,\displaystyle\mathscr{E}_{1}(\overline{\bf u}_{m},\overline{\pi}_{m})+\|\overline{\bf f}_{m}\|^{3}_{L^{p}(\mathcal{I}_{1};L^{p}(\mathcal{B}_{1}^{+}))}\to 0,\quad\ m\to\infty, (60)
τ(𝐮¯m,π¯m)>12λm3.\displaystyle\mathscr{E}_{\tau}(\overline{\bf u}_{m},\overline{\pi}_{m})>\tfrac{1}{2}\lambda^{3}_{m}. (61)

We obtain from (59)

φmφinW1,(2),φ21/p,p(2)(δ),φWy1,δ.\displaystyle\varphi_{m}\rightharpoonup^{\ast}\varphi\quad\text{in}\quad W^{1,\infty}(\mathbb{R}^{2}),\quad\varphi\in\mathcal{M}^{2-1/p,p}(\mathbb{R}^{2})(\delta),\quad\|\varphi\|_{W^{1,\infty}_{y}}\leq\delta. (62)

Hence (12) and (13) yields

𝚽m𝚽inW1,(3),𝚽2,p(3),\displaystyle{\boldsymbol{\Phi}}_{m}\rightharpoonup^{\ast}{\boldsymbol{\Phi}}\quad\text{in}\quad W^{1,\infty}(\mathbb{R}^{3}),\quad{\boldsymbol{\Phi}}\in\mathcal{M}^{2,p}(\mathbb{R}^{3}), (63)

where 𝚽m{\boldsymbol{\Phi}}_{m} is defined in accordance with (14). Finally, (16) and (17) imply

𝚿m𝚿inW1,(3),𝚿2,p(3),\displaystyle{\boldsymbol{\Psi}}_{m}\rightharpoonup^{\ast}{\boldsymbol{\Psi}}\quad\text{in}\quad W^{1,\infty}(\mathbb{R}^{3}),\quad{\boldsymbol{\Psi}}\in\mathcal{M}^{2,p}(\mathbb{R}^{3}), (64)

where 𝚿m{\boldsymbol{\Psi}}_{m} is the inverse of 𝚽m{\boldsymbol{\Phi}}_{m}. We define in 1×1+\mathcal{I}_{1}\times\mathcal{B}^{+}_{1}

𝐯¯m\displaystyle\overline{\bf v}_{m} :=1λm𝐮¯m,𝔮¯m:=1λm[π¯m(π¯m)1+],𝐠¯m:=1λm𝐟¯.\displaystyle:=\frac{1}{\lambda_{m}}\overline{\bf u}_{m},\quad\overline{\mathfrak{q}}_{m}:=\frac{1}{\lambda_{m}}\big{[}\overline{\pi}_{m}-(\overline{\pi}_{m})_{\mathcal{B}_{1}^{+}}\big{]},\quad\overline{\bf g}_{m}:=\frac{1}{\lambda_{m}}\overline{\bf f}.

We get from (60) the relation

𝒬1+|𝐯¯m|3dzdσ+(𝒬1+|𝔮¯m|5/3dzdσ)95+(𝒬1+|𝐠¯m|pdzdσ)3p=1,\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{1}}|\overline{\bf v}_{m}|^{3}\,\mathrm{d}z\,\mathrm{d}\sigma+\bigg{(}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{1}}|\overline{\mathfrak{q}}_{m}|^{5/3}\,\mathrm{d}z\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}}+\bigg{(}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{1}}|\overline{\bf g}_{m}|^{p}\,\mathrm{d}z\,\mathrm{d}\sigma\bigg{)}^{\frac{3}{p}}=1, (65)

such that, after passing to a subsequence,

𝔮¯m\displaystyle\overline{\mathfrak{q}}_{m} 𝔮¯inL5/3(1;L5/3(1+)),\displaystyle\rightharpoonup\overline{\mathfrak{q}}\quad\text{in}\quad L^{5/3}(\mathcal{I}_{1};L^{5/3}(\mathcal{B}^{+}_{1})), (66)
𝐯¯m\displaystyle\overline{\bf v}_{m} 𝐯¯inL3(1;L3(1+)),\displaystyle\rightharpoonup\overline{\bf v}\quad\text{in}\quad L^{3}(\mathcal{I}_{1};L^{3}(\mathcal{B}^{+}_{1})), (67)
𝐠¯m\displaystyle\overline{\bf g}_{m} 𝐠¯inLp(1;Lp(1+)).\displaystyle\rightharpoonup\overline{\bf g}\quad\text{in}\quad L^{p}(\mathcal{I}_{1};L^{p}(\mathcal{B}^{+}_{1})). (68)

On the other hand, (61) reads after scaling

𝒬τ+|𝐯¯m|3dzdσ\displaystyle\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}_{\tau}^{+}}|\overline{\bf v}_{m}|^{3}\,\mathrm{d}z\,\mathrm{d}\sigma +τ3(1τ+|𝔮¯m(𝔮¯m)τ+|5/3dzdσ)95>Cτ2α.\displaystyle+\tau^{3}\bigg{(}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{I}_{1}}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{B}_{\tau}^{+}}|\overline{\mathfrak{q}}_{m}-(\overline{\mathfrak{q}}_{m})_{\mathcal{B}_{\tau}^{+}}|^{5/3}\,\mathrm{d}z\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}}>C_{\ast}\tau^{2\alpha}. (69)

In order to proceed we use a scaled version of the equation which reads as

Jφmt𝐯¯m+λm(𝐁φm𝐯¯m)𝐯¯m+div(𝐁φm𝔮¯m)div(𝐀φm𝐯¯m)=Jφm𝐠¯m,𝐁φm:𝐯¯m=0,𝐯¯m|1+=0.\displaystyle\begin{aligned} J_{\varphi_{m}}\partial_{t}\overline{\bf v}_{m}+\lambda_{m}({\bf B}_{\varphi_{m}}\nabla\overline{\bf v}_{m})\overline{\bf v}_{m}+\operatorname{div}\big{(}{\bf B}_{\varphi_{m}}\overline{\mathfrak{q}}_{m}\big{)}-\operatorname{div}\big{(}{\bf A}_{\varphi_{m}}\nabla\overline{\bf v}_{m}\big{)}&=J_{\varphi_{m}}\overline{\bf g}_{m},\\ {\bf B}_{\varphi_{m}}:\nabla{\overline{\bf v}}_{m}=0,\quad\overline{\bf v}_{m}|_{\mathcal{B}^{+}_{1}\cap\partial\mathbb{H}}&=0.\end{aligned} (70)

By (45) (replacing ζ\zeta by ζ2\zeta^{2}) we have

1+12Jφmζ2\displaystyle\int_{\mathcal{B}^{+}_{1}}\frac{1}{2}J_{\varphi_{m}}\zeta^{2} |𝐯¯m(t)|2dx+0t1+ζ2𝐀φm𝐯¯m:𝐯¯mdxdσ\displaystyle\big{|}\overline{\bf v}_{m}(t)\big{|}^{2}\,\mathrm{d}x+\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}\zeta^{2}{\bf A}_{\varphi_{m}}\nabla\overline{\bf v}_{m}:\nabla\overline{\bf v}_{m}\,\mathrm{d}x\,\mathrm{d}\sigma
0t1+12Jφm|𝐯¯m|2tζ2dxdσ+0t1+12Jφm|𝐯¯m|2Δ𝚿m𝚽mζ2dxdσ\displaystyle\leq\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}\frac{1}{2}J_{\varphi_{m}}|\overline{\bf v}_{m}|^{2}\partial_{t}\zeta^{2}\,\mathrm{d}x\,\mathrm{d}\sigma+\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}\frac{1}{2}J_{\varphi_{m}}|\overline{\bf v}_{m}|^{2}\Delta{{\boldsymbol{\Psi}}_{m}}\circ{\boldsymbol{\Phi}}_{m}\cdot\nabla\zeta^{2}\,\mathrm{d}x\,\mathrm{d}\sigma
+0t1+12|𝐯¯m|2𝐀φm:2ζ2dxdσ+0t1+12(λm|𝐯¯m|2+2𝔮¯m)𝐯¯m𝐁φmζ2)dxdσ\displaystyle+\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}\frac{1}{2}|\overline{\bf v}_{m}|^{2}{\bf A}_{\varphi_{m}}:\nabla^{2}\zeta^{2}\,\mathrm{d}x\,\mathrm{d}\sigma+\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}\frac{1}{2}\big{(}\lambda_{m}|\overline{\bf v}_{m}|^{2}+2\overline{\mathfrak{q}}_{m}\big{)}\overline{\bf v}_{m}\cdot{\bf B}_{\varphi_{m}}\nabla\zeta^{2}\Big{)}\,\mathrm{d}x\,\mathrm{d}\sigma
+0tΩJφmζ2𝐠¯m𝐯¯mdxdσ\displaystyle+\int_{0}^{t}\int_{\Omega}J_{\varphi_{m}}\zeta^{2}\overline{\bf g}_{m}\cdot\overline{\bf v}_{m}\,\mathrm{d}x\,\mathrm{d}\sigma

for all non-negative ζCc(𝒬1)\zeta\in C^{\infty}_{c}(\mathcal{Q}_{1}). Further, note that by (63) and (64) we know that JφmJ_{\varphi_{m}}, 𝐀φm{\bf A}_{\varphi_{m}} and 𝐁φm{\bf B}_{\varphi_{m}} are uniformly bounded. Hence the first, third, fourth and fifth term are uniformly bounded by (65). The second term is more delicate and we need to employ Sobolev mutltipliers (see Section 2.3 for a brief introduction). We obtain by (63), (64) and (65)

0t1+\displaystyle\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}} 12Jφm|𝐯¯m|2Δ𝚿m𝚽mζ2dxdσ\displaystyle\frac{1}{2}J_{\varphi_{m}}|\overline{{\bf v}}_{m}|^{2}\Delta{{\boldsymbol{\Psi}}_{m}}\circ{\boldsymbol{\Phi}}_{m}\cdot\nabla\zeta^{2}\,\mathrm{d}x\,\mathrm{d}\sigma
k=130t1+|𝐯¯m||ζ𝐯¯mΔ𝚿mk𝚽m|dxdσ\displaystyle\lesssim\sum_{k=1}^{3}\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}|\overline{{\bf v}}_{m}||\zeta\overline{{\bf v}}_{m}\Delta{\boldsymbol{\Psi}}_{m}^{k}\circ{\boldsymbol{\Phi}}_{m}|\,\mathrm{d}x\,\mathrm{d}\sigma
0t1+|𝐯¯m|3dxdσ+k=130t1+|ζ𝐯¯mΔ𝚿mk𝚽m|3/2dxdσ\displaystyle\lesssim\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}|\overline{{\bf v}}_{m}|^{3}\,\mathrm{d}x\,\mathrm{d}\sigma+\sum_{k=1}^{3}\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}|\zeta\overline{{\bf v}}_{m}\Delta{\boldsymbol{\Psi}}_{m}^{k}\circ{\boldsymbol{\Phi}}_{m}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma
1+k=130t𝚽m(1+)|(ζ𝐯¯m)𝚿mΔ𝚿mk|3/2dxdσ=:1+(I)m.\displaystyle\lesssim 1+\sum_{k=1}^{3}\int_{0}^{t}\int_{{\boldsymbol{\Phi}}_{m}(\mathcal{B}^{+}_{1})}|(\zeta\overline{{\bf v}}_{m})\circ{\boldsymbol{\Psi}}_{m}\Delta{\boldsymbol{\Psi}}_{m}^{k}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma=:1+(I)_{m}.

The remaining integral (I)m(I)_{m} can be split into

(I)m\displaystyle(I)_{m} k=130t𝚽m(1+)|((ζ𝐯¯m)𝚿m)𝚿mk|3/2dxdσ\displaystyle\lesssim\sum_{k=1}^{3}\int_{0}^{t}\int_{{\boldsymbol{\Phi}}_{m}(\mathcal{B}^{+}_{1})}|\nabla((\zeta\overline{{\bf v}}_{m})\circ{\boldsymbol{\Psi}}_{m})\nabla{\boldsymbol{\Psi}}_{m}^{k}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma
+k=130t𝚽m(1+)|div((ζ𝐯¯m)𝚿m𝚿mk)|3/2dxdσ=:(I)m1+(I)m2.\displaystyle+\sum_{k=1}^{3}\int_{0}^{t}\int_{{\boldsymbol{\Phi}}_{m}(\mathcal{B}^{+}_{1})}|\operatorname{div}((\zeta\overline{{\bf v}}_{m})\circ{\boldsymbol{\Psi}}_{m}\otimes\nabla{\boldsymbol{\Psi}}_{m}^{k})|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma=:(I)_{m}^{1}+(I)_{m}^{2}.

By (63) and (64) we have for κ>0\kappa>0 arbitrary

(I)m1\displaystyle(I)_{m}^{1} k=130t𝚽m(1+)|(ζ𝚿m)𝐯¯m𝚿m𝚿mk|3/2dxdσ\displaystyle\lesssim\sum_{k=1}^{3}\int_{0}^{t}\int_{{\boldsymbol{\Phi}}_{m}(\mathcal{B}^{+}_{1})}|\nabla(\zeta\circ{\boldsymbol{\Psi}}_{m})\overline{{\bf v}}_{m}\circ{\boldsymbol{\Psi}}_{m}\nabla{\boldsymbol{\Psi}}_{m}^{k}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma
+k=130t𝚽m(1+)|ζ𝚿m(𝐯¯m𝚿m)𝚿mk|3/2dxdσ\displaystyle+\sum_{k=1}^{3}\int_{0}^{t}\int_{{\boldsymbol{\Phi}}_{m}(\mathcal{B}^{+}_{1})}|\zeta\circ{\boldsymbol{\Psi}}_{m}\nabla(\overline{{\bf v}}_{m}\circ{\boldsymbol{\Psi}}_{m})\nabla{\boldsymbol{\Psi}}_{m}^{k}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma
0t𝚽m(1+)|(ζ𝚿m)𝐯¯m𝚿m|3/2dxdσ\displaystyle\lesssim\int_{0}^{t}\int_{{\boldsymbol{\Phi}}_{m}(\mathcal{B}^{+}_{1})}|\nabla(\zeta\circ{\boldsymbol{\Psi}}_{m})\overline{{\bf v}}_{m}\circ{\boldsymbol{\Psi}}_{m}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma
+0t𝚽m(1+)|ζ𝚿m(𝐯¯m𝚿m)|3/2dxdσ\displaystyle+\int_{0}^{t}\int_{{\boldsymbol{\Phi}}_{m}(\mathcal{B}^{+}_{1})}|\zeta\circ{\boldsymbol{\Psi}}_{m}\nabla(\overline{{\bf v}}_{m}\circ{\boldsymbol{\Psi}}_{m})|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma
0t1+|𝐯¯m|3/2dxdσ+0t1+|ζ𝐯¯m|3/2dxdσ\displaystyle\lesssim\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}|\overline{{\bf v}}_{m}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma+\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}|\zeta\nabla\overline{{\bf v}}_{m}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma
1+κ0t1+ζ2|𝐯¯m|2dxdσ\displaystyle\lesssim 1+\kappa\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}\zeta^{2}|\nabla\overline{{\bf v}}_{m}|^{2}\,\mathrm{d}x\,\mathrm{d}\sigma (71)

using also (65) in the last step. Moreover, it holds by (64) and (9)555Note that the assumption p3/2p\geq 3/2 would have been sufficient here, see [19, Proposition 3.5.3.] for inclusions of the spaces s,p\mathcal{M}^{s,p}.

(I)m2k=130t(ζ𝐯¯m)𝚿mW1,3/2(𝚽m(1+))3/2𝚿mk2,3/2(𝚽m(1+))3/2dσk=130t|ζ𝐯¯mW1,3/2(1+)3/2dσ0t1+|𝐯¯m|3/2dxdσ+0t1+|ζ𝐯¯m|3/2dxdσ1+κ0t1+ζ2|𝐯¯m|2dxdσ.\displaystyle\begin{aligned} (I)_{m}^{2}&\lesssim\sum_{k=1}^{3}\int_{0}^{t}\|(\zeta\overline{{\bf v}}_{m})\circ{\boldsymbol{\Psi}}_{m}\|_{W^{1,3/2}({\boldsymbol{\Phi}}_{m}(\mathcal{B}^{+}_{1}))}^{3/2}\|{\boldsymbol{\Psi}}_{m}^{k}\|_{\mathcal{M}^{2,3/2}({\boldsymbol{\Phi}}_{m}(\mathcal{B}^{+}_{1}))}^{3/2}\,\mathrm{d}\sigma\\ &\lesssim\sum_{k=1}^{3}\int_{0}^{t}|\|\zeta\overline{{\bf v}}_{m}\|_{W^{1,3/2}(\mathcal{B}_{1}^{+})}^{3/2}\,\mathrm{d}\sigma\\ &\lesssim\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}|\overline{{\bf v}}_{m}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma+\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}|\zeta\nabla\overline{{\bf v}}_{m}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma\\ &\lesssim 1+\kappa\int_{0}^{t}\int_{\mathcal{B}^{+}_{1}}\zeta^{2}|\nabla\overline{{\bf v}}_{m}|^{2}\,\mathrm{d}x\,\mathrm{d}\sigma.\end{aligned} (72)

Let us choose ζ\zeta such that ζ=1\zeta=1 in 𝒬3/4\mathcal{Q}_{3/4}. Finally, since 𝐀φm{\bf A}_{\varphi_{m}} is elliptic uniformly in mm and JφmJ_{\varphi_{m}} strictly positive by (16), we conclude

𝐯¯mL(3/4;L2(3/4+))L2(3/4;W1,2(3/4+))\displaystyle\overline{{\bf v}}_{m}\in L^{\infty}\big{(}\mathcal{I}_{3/4};L^{2}\big{(}\mathcal{B}^{+}_{3/4}\big{)}\big{)}\cap L^{2}\big{(}\mathcal{I}_{3/4};W^{1,2}\big{(}\mathcal{B}_{3/4}^{+}\big{)}\big{)} (73)

uniformly in mm choosing κ\kappa small enough. By Sobolev’s inequality we have for ϕCc(𝒬3/4+){\boldsymbol{\phi}}\in C^{\infty}_{c}(\mathcal{Q}_{3/4}^{+})

𝒬1+\displaystyle\int_{\mathcal{Q}^{+}_{1}} t𝐯¯mϕdzdσc(𝐯¯mLz2𝐯¯mLz2+𝔮¯mLz3/2+𝐠¯mLz2)ϕWz2,2.\displaystyle\partial_{t}\overline{{\bf v}}_{m}\cdot{\boldsymbol{\phi}}\,\mathrm{d}z\,\mathrm{d}\sigma\leq\,c\,\Big{(}\|\overline{{\bf v}}_{m}\|_{L^{2}_{z}}\|\nabla\overline{{\bf v}}_{m}\|_{L^{2}_{z}}+\|\overline{\mathfrak{q}}_{m}\|_{L^{3/2}_{z}}+\|\overline{\bf g}_{m}\|_{L^{2}_{z}}\Big{)}\|{\boldsymbol{\phi}}\|_{W^{2,2}_{z}}.

This shows by (61) and (73) the boundedness of

t𝐯¯mL2(3/4;W2,2(3/4+)).\displaystyle\partial_{t}\overline{{\bf v}}_{m}\in L^{2}(\mathcal{I}_{3/4};W^{-2,2}(\mathcal{B}^{+}_{3/4})). (74)

After passing to suitable subsequences we obtain

t𝐯¯m\displaystyle\partial_{t}\overline{{\bf v}}_{m} t𝐯¯inL2(3/4;W2,2(3/4+)).\displaystyle\rightharpoonup\partial_{t}\overline{{\bf v}}\quad\text{in}\quad L^{2}\big{(}\mathcal{I}_{3/4};W^{-2,2}\big{(}\mathcal{B}^{+}_{3/4}\big{)}\big{)}. (75)

Now, (73) and (74) imply

𝐯¯m\displaystyle\overline{\bf v}_{m} 𝐯¯inL3(3/4;L3(3/4+))\displaystyle\rightarrow\overline{\bf v}\quad\text{in}\quad L^{3}\big{(}\mathcal{I}_{3/4};L^{3}\big{(}\mathcal{B}_{3/4}^{+}\big{)}\big{)} (76)

by the Aubin-Lions compactness theorem. Recalling the definitions of 𝐀φm{\bf A}_{\varphi_{m}} and 𝐁φm{\bf B}_{\varphi_{m}} from (42) we are able to pass to the limit in (70) using the convergences (63), (64), (66), (67) (73) and (75). We obtain (in the sense of distributions on 𝒬3/4+\mathcal{Q}^{+}_{3/4})

Jφt𝐯¯+div(𝐁φ𝔮¯)div(𝐀φ𝐯¯)=Jφ𝐠¯,𝐁φ:𝐯¯=0,𝐯¯|3/4+=0,\displaystyle\begin{aligned} J_{\varphi}\partial_{t}\overline{\bf v}+\operatorname{div}\big{(}{\bf B}_{\varphi}\overline{\mathfrak{q}}\big{)}-\operatorname{div}\big{(}{\bf A}_{\varphi}\nabla\overline{\bf v}\big{)}&=J_{\varphi}\overline{\bf g},\\ {\bf B}_{\varphi}^{\top}:\nabla{\overline{\bf v}}=0,\quad\overline{\bf v}|_{\mathcal{B}^{+}_{3/4}\cap\partial\mathbb{H}}&=0,\end{aligned} (77)

recalling that the comvective term disappears as λm0\lambda_{m}\rightarrow 0. Note that since p>3p>3 relation (8) yields boundedness of 𝚿m{\boldsymbol{\Psi}}_{m} and 𝚽m{\boldsymbol{\Phi}}_{m} in W2,pW^{2,p} such that 𝚿m\nabla{\boldsymbol{\Psi}}_{m} and 𝚽m\nabla{\boldsymbol{\Phi}}_{m} are pre-compact.

We are now fully prepared to lead (69) to a contradiction applying the regularity theory developed in Section 4 to (70) and (77). We infer from Lemma 4.3 applied to (77), using also (62), that

2𝐯¯L5/3(1/2;Lp(1/2+))\displaystyle\|\nabla^{2}\overline{\bf v}\|_{L^{5/3}(\mathcal{I}_{1/2};L^{p}(\mathcal{B}^{+}_{1/2}))} +t𝐯¯L5/3(1/2;Lp(1/2+))\displaystyle+\|\partial_{t}\overline{\bf v}\|_{L^{5/3}(\mathcal{I}_{1/2};L^{p}(\mathcal{B}^{+}_{1/2}))}
𝐯¯L5/3(3/4;L5/3(3/4+))+𝔮¯L5/3(3/4;L5/3(3/4+))+𝐠¯L5/3(3/4;Lp(3/4+)),\displaystyle\lesssim\|\nabla\overline{\bf v}\|_{L^{5/3}(\mathcal{I}_{3/4};L^{5/3}(\mathcal{B}^{+}_{3/4}))}+\|\overline{\mathfrak{q}}\|_{L^{5/3}(\mathcal{I}_{3/4};L^{5/3}(\mathcal{B}^{+}_{3/4}))}+\|\overline{\bf g}\|_{L^{5/3}(\mathcal{I}_{3/4};L^{p}(\mathcal{B}^{+}_{3/4}))},

where the right-hand side is finite on account of (66), (68) and (73). We deduce from parabolic embeddings that 𝐯¯\overline{\bf v} belongs to the class Cparaβ/2,β(𝒬¯3/4+)C^{\beta/2,\beta}_{\mathrm{para}}(\overline{\mathcal{Q}}^{+}_{3/4}), where β=4/53/p>α\beta=4/5-3/p>\alpha. Note that β>0\beta>0 for p>154p>\frac{15}{4}. This, (76) and 𝐯¯|3/4+=0\overline{\bf v}|_{\mathcal{B}^{+}_{3/4}\cap\partial\mathbb{H}}=0 prove

limm𝒬τ+|𝐯¯m|3dzdσ=𝒬τ+|𝐯¯|3dzdσC𝐯¯τ3α\displaystyle\lim_{m\rightarrow\infty}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{\tau}}|\overline{\bf v}_{m}|^{3}\,\mathrm{d}z\,\mathrm{d}\sigma=\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{\tau}}|\overline{\bf v}|^{3}\,\mathrm{d}z\,\mathrm{d}\sigma\leq C_{\overline{\bf v}}\tau^{3\alpha} (78)

for all τ<12\tau<\frac{1}{2} for some constant C𝐯¯>0C_{\overline{\bf v}}>0. Let us consider the unique solution (𝐯~m,𝔮~m)(\tilde{\bf v}_{m},\tilde{\mathfrak{q}}_{m}) to

Jφmt𝐯~m+div(𝐁φm𝔮~m)div(𝐀φm𝐯~m)=λm(𝐁φm𝐯¯m)𝐯¯m,𝐁φm:𝐯~m=0,𝐯~m|3/4+=0,𝐯¯m(3/4,)=0,\displaystyle\begin{aligned} J_{\varphi_{m}}\partial_{t}\tilde{\bf v}_{m}+\operatorname{div}\big{(}{\bf B}_{\varphi_{m}}\tilde{\mathfrak{q}}_{m}\big{)}-\operatorname{div}\big{(}{\bf A}_{\varphi_{m}}\nabla\tilde{\bf v}_{m}\big{)}&=-\lambda_{m}({\bf B}_{\varphi_{m}}\nabla\overline{\bf v}_{m})\overline{\bf v}_{m},\\ {\bf B}_{\varphi_{m}}^{\top}:\nabla\tilde{{\bf v}}_{m}=0,\quad\tilde{\bf v}_{m}|_{\mathcal{B}^{+}_{3/4}\cap\partial\mathbb{H}}&=0,\quad\overline{{\bf v}}_{m}(-3/4,\cdot)=0,\end{aligned} (79)

where the first equation is understood in the sense of distributions on 𝒬3/4+\mathcal{Q}^{+}_{3/4}. By the regularity theory for the perturbed Stokes system established in Lemma 4.2 (which applies on account of (62)) we have in 𝒬3/4+\mathcal{Q}^{+}_{3/4}

2𝐯~mLt5/3Lx15/14+𝔮~mLt5/3Lx15/14λm(𝐁φm𝐯¯m)𝐯¯mLt5/3Lx15/140,m,\displaystyle\begin{aligned} \|\nabla^{2}\tilde{\bf v}_{m}\|_{L^{5/3}_{t}L^{{15/14}}_{x}}&+\|\nabla\tilde{\mathfrak{q}}_{m}\|_{L^{5/3}_{t}L^{{15/14}}_{x}}\\ &\lesssim\lambda_{m}\|({\bf B}_{\varphi_{m}}\nabla\overline{\bf v}_{m})\overline{\bf v}_{m}\|_{L^{5/3}_{t}L^{{15/14}}_{x}}\longrightarrow 0,\quad m\rightarrow\infty,\end{aligned} (80)

using (59) and (73). Now we consider the difference 𝐯~m=𝐯~m𝐯¯m\tilde{\bf v}^{\ast}_{m}=\tilde{\bf v}_{m}-\overline{\bf v}_{m}, 𝔮~m=𝔮~m𝔮¯m\tilde{\mathfrak{q}}_{m}^{\ast}=\tilde{\mathfrak{q}}_{m}-\overline{\mathfrak{q}}_{m} which solves a Stokes system with right-hand side Jφm𝐠¯mJ_{\varphi_{m}}\overline{\bf g}_{m} in 𝒬3/4+\mathcal{Q}_{3/4}^{+} and hence satisfies

𝔮~mLt5/3Lxp𝐠¯mLt5/3Lxp+𝐯~mLt5/3Lx15/14+𝔮~m(𝔮~m)3/4+L5/3L15/14\displaystyle\|\nabla\tilde{\mathfrak{q}}^{\ast}_{m}\|_{L^{5/3}_{t}L^{p}_{x}}\lesssim\|\overline{\bf g}_{m}\|_{L^{5/3}_{t}L^{p}_{x}}+\|\nabla\tilde{\bf v}^{\ast}_{m}\|_{L^{5/3}_{t}L^{{15/14}}_{x}}+\|\tilde{\mathfrak{q}}^{\ast}_{m}-(\tilde{\mathfrak{q}}_{m}^{\ast})_{\mathcal{B}^{+}_{3/4}}\|_{L^{5/3}L^{{15/14}}}

by Lemma 4.3 using also (62). Here the norm on the left-hand side is taken over 𝒬1/2+\mathcal{Q}_{1/2}^{+} and the one on the right-hand side over 𝒬3/4+\mathcal{Q}_{3/4}^{+}. The first term on the right-hand side is bounded by (65) and the second one by (73) and (80) (recall that 𝐯~m\tilde{\bf v}_{m} has zero boundary conditions). For the third one we have

𝔮~m(𝔮~m)3/4+Lx5/3Lx15/14\displaystyle\|\tilde{\mathfrak{q}}^{\ast}_{m}-(\tilde{\mathfrak{q}}^{\ast}_{m})_{\mathcal{B}^{+}_{3/4}}\|_{L^{5/3}_{x}L^{{15/14}}_{x}} 𝔮¯m(𝔮¯m)3/4+Lt5/3Lx15/14+𝔮~m(𝔮~m)3/4+Lt5/3Lx15/14\displaystyle\lesssim\|\overline{\mathfrak{q}}_{m}-(\overline{\mathfrak{q}}_{m})_{\mathcal{B}^{+}_{3/4}}\|_{L^{5/3}_{t}L^{{15/14}}_{x}}+\|\tilde{\mathfrak{q}}_{m}-(\tilde{\mathfrak{q}}_{m})_{\mathcal{B}^{+}_{3/4}}\|_{L^{5/3}_{t}L^{{15/14}}_{x}}
𝔮¯mLt5/3Lx15/14+𝔮~mLt5/3Lx15/14,\displaystyle\lesssim\|\overline{\mathfrak{q}}_{m}\|_{L^{5/3}_{t}L^{{15/14}}_{x}}+\|\nabla\tilde{\mathfrak{q}}_{m}\|_{L^{5/3}_{t}L^{{15/14}}_{x}},

which is bounded by (66) and (80). We conclude, that 𝔮~mLt5/3Lxp\|\nabla\tilde{\mathfrak{q}}^{\ast}_{m}\|_{L^{5/3}_{t}L^{p}_{x}} (with norm taken over 𝒬1/2+\mathcal{Q}_{1/2}^{+}) is bounded which yields

τ3(ττ+|𝔮~m(𝔮~m)τ+|5/3dzdσ)95\displaystyle\tau^{3}\bigg{(}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{I}_{\tau}}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{B}^{+}_{\tau}}|\tilde{\mathfrak{q}}^{\ast}_{m}-(\tilde{\mathfrak{q}}_{m}^{\ast})_{\mathcal{B}^{+}_{\tau}}|^{5/3}\,\mathrm{d}z\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}} τ3(τττ+|𝔮~m|5/3dzdσ)95\displaystyle\lesssim\tau^{-3}\bigg{(}\int_{-\tau}^{\tau}\int_{\mathcal{B}^{+}_{\tau}}|\nabla\tilde{\mathfrak{q}}^{\ast}_{m}|^{5/3}\,\mathrm{d}z\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}}
τ3(τττ35p(τ+|𝔮~m|pdz)53pdσ)95\displaystyle\leq\tau^{-3}\bigg{(}\int_{-\tau}^{\tau}\tau^{3-\frac{5}{p}}\bigg{(}\int_{\mathcal{B}^{+}_{\tau}}|\nabla\tilde{\mathfrak{q}}^{\ast}_{m}|^{p}\,\mathrm{d}z\bigg{)}^{\frac{5}{3p}}\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}}
τ2α,\displaystyle\lesssim\tau^{2\alpha},

where α=3(2532p)\alpha=3\big{(}\frac{2}{5}-\frac{3}{2p}\big{)}. Combining this with (80) and Sobolev’s embedding Wx1,15/14Lx5/3W^{1,15/14}_{x}\hookrightarrow L^{5/3}_{x} shows

lim supmτ3(\displaystyle\limsup_{m}\tau^{3}\bigg{(} τ𝒬τ+|𝔮¯m(𝔮¯m)τ+|5/3dzdσ)95\displaystyle\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{I}_{\tau}}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{\tau}}|\overline{\mathfrak{q}}_{m}-(\overline{\mathfrak{q}}_{m})_{\mathcal{B}^{+}_{\tau}}|^{5/3}\,\mathrm{d}z\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}}
=lim supmτ3(ττ+|𝔮~m(𝔮~m)τ+|5/3dzdσ)95C𝔮¯τ2α\displaystyle=\limsup_{m}\tau^{3}\bigg{(}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{I}_{\tau}}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{B}_{\tau}^{+}}|\tilde{\mathfrak{q}}^{\ast}_{m}-(\tilde{\mathfrak{q}}^{\ast}_{m})_{\mathcal{B}^{+}_{\tau}}|^{5/3}\,\mathrm{d}z\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}}\leq\,C_{\overline{\mathfrak{q}}}\tau^{2\alpha}

for some constant C𝔮¯C_{\overline{\mathfrak{q}}}. This, together with (78), contradicts (69) if we choose C>C𝐯¯+C𝔮¯C_{\ast}>C_{\overline{\bf v}}+C_{\overline{\mathfrak{q}}}. ∎

Proof of Theorem 2.6.

By assumption we have

r2t0r2t0+r2Ωr(x0)|𝐮|3dxdt+(r5/3t0r2t0+r2Ωr(x0)|π|5/3dxdt)95<ε0.\displaystyle r^{-2}\int_{t_{0}-r^{2}}^{t_{0}+r^{2}}\int_{\Omega\cap\mathcal{B}_{r}(x_{0})}|{\bf u}|^{3}\,\mathrm{d}x\,\mathrm{d}t+\bigg{(}r^{-5/3}\int_{t_{0}-r^{2}}^{t_{0}+r^{2}}\int_{\Omega\cap\mathcal{B}_{r}(x_{0})}|\pi|^{5/3}\,\mathrm{d}x\,\mathrm{d}t\bigg{)}^{\frac{9}{5}}<\varepsilon_{0}.

After rotation and translation of the coordinate system we can assume that (t0,x0)=(0,0)(t_{0},x_{0})=(0,0). With the mapping 𝚽{\boldsymbol{\Phi}} from (14) we turn to the functions π¯=π𝚽\overline{\pi}=\pi\circ{\boldsymbol{\Phi}} and 𝐮¯=𝐮𝚽\overline{{\bf u}}={\bf u}\circ{\boldsymbol{\Phi}} which solve the perturbed system (43) in I×λ+I\times\mathcal{B}^{+}_{\lambda\mathfrak{R}}. Here \mathfrak{R} is given in Remark 2.3 and λ\lambda is chosen such that 𝚽(λ+)ΩB{\boldsymbol{\Phi}}(\mathcal{B}^{+}_{\lambda\mathfrak{R}})\subset\Omega\cap B_{\mathfrak{R}} (recall that 𝚽{\boldsymbol{\Phi}} is Lipschitz). We obtain with R:=λrR:=\lambda r

R3Rt0,x0\displaystyle R^{3}\mathscr{E}_{R}^{t_{0},x_{0}} (𝐮¯,π¯)R2t0R2t0+R2R+|𝐮¯|3dxdt+R3(t0R2t0+R2R+|π¯|5/3dxdt)95\displaystyle(\overline{\bf u},\overline{\pi})\lesssim R^{-2}\int_{t_{0}-R^{2}}^{t_{0}+R^{2}}\int_{\mathcal{B}^{+}_{R}}|\overline{\bf u}|^{3}\,\mathrm{d}x\,\mathrm{d}t+R^{-3}\bigg{(}\int_{t_{0}-R^{2}}^{t_{0}+R^{2}}\int_{\mathcal{B}_{R}^{+}}|\overline{\pi}|^{5/3}\,\mathrm{d}x\,\mathrm{d}t\bigg{)}^{\frac{9}{5}}
r2t0r2t0+r2Ωr(x0)|𝐮|3dxdt+(r5/3t0r2t0+r2Ωr(x0)|π|5/3dxdt)95<ε0.\displaystyle\lesssim r^{-2}\int_{t_{0}-r^{2}}^{t_{0}+r^{2}}\int_{\Omega\cap\mathcal{B}_{r}(x_{0})}|{\bf u}|^{3}\,\mathrm{d}x\,\mathrm{d}t+\bigg{(}r^{-5/3}\int_{t_{0}-r^{2}}^{t_{0}+r^{2}}\int_{\Omega\cap\mathcal{B}_{r}(x_{0})}|\pi|^{5/3}\,\mathrm{d}x\,\mathrm{d}t\bigg{)}^{\frac{9}{5}}<\varepsilon_{0}.

Furthermore, it holds

R9(t0R2t0+R2R+|𝐟¯|pdxdt)3p=(R3p5t0R2t0+R2R+|𝐟¯|pdxdt)3p<ε0\displaystyle R^{9}\bigg{(}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{t_{0}-R^{2}}^{t_{0}+R^{2}}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{B}_{R}^{+}}|\overline{\bf f}|^{p}\,\mathrm{d}x\,\mathrm{d}t\bigg{)}^{\frac{3}{p}}=\bigg{(}R^{3p-5}\int_{t_{0}-R^{2}}^{t_{0}+R^{2}}\int_{\mathcal{B}_{R}^{+}}|\overline{\bf f}|^{p}\,\mathrm{d}x\,\mathrm{d}t\bigg{)}^{\frac{3}{p}}<\varepsilon_{0}

provided we choose RR small enough. If ε0\varepsilon_{0} is small enough we obtain

R3Rt0,x0(𝐮¯,π¯)+R9(𝒬R+|𝐟¯|pdydσ)3pε,R^{3}\mathscr{E}_{R}^{t_{0},x_{0}}(\overline{\bf u},\overline{\pi})+R^{9}\bigg{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\mathcal{Q}^{+}_{R}}|\overline{\bf f}|^{p}\,\mathrm{d}y\,\mathrm{d}\sigma\bigg{)}^{\frac{3}{p}}\leq\varepsilon,

such that a scaled version of Lemma 5.1 applies (note that φ\varphi must be scaled to 1Rφ(R)\frac{1}{R}\varphi(R\cdot) such that its multiplier norm and Lipschitz constant are still bounded by δ\delta as long as R1R\leq 1). We conclude that

R3τRt0,x0(𝐮¯,π¯)Cτ2α(R3Rt0,x0(𝐮¯,π¯)+R9(𝒬R+|𝐟¯|pdydσ)3p).R^{3}\mathscr{E}^{t_{0},x_{0}}_{\tau R}(\overline{\bf u},\overline{\pi})\leq C_{\ast}\tau^{2\alpha}\bigg{(}R^{3}\mathscr{E}_{R}^{t_{0},x_{0}}(\overline{\bf u},\overline{\pi})+R^{9}\bigg{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\mathcal{Q}^{+}_{R}}|\overline{\bf f}|^{p}\,\mathrm{d}y\,\mathrm{d}\sigma\bigg{)}^{\frac{3}{p}}\bigg{)}.

It is standard to iterate this inequality (see, e.g., [11, Prop. 2.5]), which yields

τkRt0,x0(𝐮¯,π¯)τ2αk.\displaystyle\mathscr{E}_{\tau^{k}R}^{t_{0},x_{0}}(\overline{\bf u},\overline{\pi})\lesssim\tau^{2\alpha k}.

We trivially obtain

~τkRt0,x0(𝐮¯,π¯)τ2αk,\displaystyle\widetilde{\mathscr{E}}_{\tau^{k}R}^{t_{0},x_{0}}(\overline{\bf u},\overline{\pi})\lesssim\tau^{2\alpha k}, (81)

where the access ~\widetilde{\mathscr{E}} is given by

~rt0,x0(𝐮¯,π¯)\displaystyle\widetilde{\mathscr{E}}_{r}^{t_{0},x_{0}}(\overline{\bf u},\overline{\pi}) :=𝒬r(t0,x0)(I×Ω)|𝐮¯(𝐮¯)𝒬r(t0,x0)(I×Ω)|3dydσ\displaystyle:=\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}_{r}(t_{0},x_{0})\cap(I\times\Omega)}|\overline{\bf u}-(\overline{\bf u})_{\mathcal{Q}_{r}(t_{0},x_{0})\cap(I\times\Omega)}|^{3}\,\mathrm{d}y\,\mathrm{d}\sigma
+r3(r(t0)r+(x0)|π¯(π¯)r(x0)Ω|5/3dydσ)95.\displaystyle+r^{3}\bigg{(}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{I}_{r}(t_{0})}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{B}^{+}_{r}(x_{0})}|\overline{\pi}-(\overline{\pi})_{\mathcal{B}_{r}(x_{0})\cap\Omega}|^{5/3}\,\mathrm{d}y\,\mathrm{d}\sigma\bigg{)}^{\frac{9}{5}}.

The decay estimate (81) holds in the centre point (t0,x0)=(0,0)(t_{0},x_{0})=(0,0) of 𝒬1+\mathcal{Q}_{1}^{+} but also for points (t0,x0)(t_{0},x_{0}) on I×ΩI\times\partial\Omega which are sufficiently close. We claim that it continues to hold in the interior of Ω\Omega. In fact, we can prove a version of Lemma 5.1 for interior points for which 𝒬R(t0,x0)𝒬+\mathcal{Q}_{R}(t_{0},x_{0})\subset\mathcal{Q}^{+}. The main difference is that one has to replace in the proof Lemmas 4.2 and 4.3 by their corresponding interior versions Lemmas 4.4 and 4.5. Combining the interior and the boundary version yields

~τkRt0,x0(𝐮,π)r2α.\displaystyle\widetilde{\mathscr{E}}_{\tau^{k}R}^{t_{0},x_{0}}({\bf u},\pi)\lesssim r^{2\alpha}.

This proves that 𝐮¯C0,α\overline{\bf u}\in C^{0,\alpha} in a neighborhood of (0,0)(0,0). Changing coordinates (that is, recalling that 𝐮=𝐮¯𝚿{\bf u}=\overline{\bf u}\circ{\boldsymbol{\Psi}} where 𝚿{\boldsymbol{\Psi}} is Lipschitz, cf. (16)) and changing the coordinate system as the case may be, proves 𝐮C0,α(𝒰¯(t0,x0)){\bf u}\in C^{0,\alpha}(\overline{\mathcal{U}}(t_{0},x_{0})) for some neighborhood 𝒰(t0,x0)\mathcal{U}(t_{0},x_{0}) of (t0,x0)(t_{0},x_{0}). ∎

6. Concluding remarks & outlook

6.1. The size of the singular set

We comment in this section on the fact that our estimate on the size of the singular set in Theorem 2.7 (dimension 5/3\leq 5/3) is weaker than that from [26] for regular boundaries. We introduce the dissipation functional

𝒟r(𝐮¯)=𝒬r+|𝐮¯|2dxdt.\displaystyle\mathscr{D}_{r}(\overline{\bf u})=\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}_{r}^{+}}|\nabla\overline{\bf u}|^{2}\,\mathrm{d}x\,\mathrm{d}t.

The following implication is proved in [26, proof of Theorem 5.1]:

supr<11r𝒬r+|𝐮¯|2dxdt<ε𝒟lim infr0r3r(𝐮¯,π¯)<ε,\displaystyle\sup_{r<1}\frac{1}{r}\int_{\mathcal{Q}_{r}^{+}}|\nabla\overline{\bf u}|^{2}\,\mathrm{d}x\,\mathrm{d}t<\varepsilon_{\mathscr{D}}\quad\Rightarrow\quad\liminf_{r\rightarrow 0}r^{3}\mathscr{E}_{r}(\overline{\bf u},\overline{\pi})<\varepsilon, (82)

where ε𝒟\varepsilon_{\mathscr{D}} is a small number. The size of the set where the first condition is violated is much smaller than the size of the set where the second one is violated (the parabolic Hausdorff-dimensions are 1 and 5/3). Let us explain the strategy to prove the implication (82). Transforming the local energy inequality to the flat geometry (as we did in (45)) and diving by rr, they proof (in our notation)

sup3r/41r3r/4+|𝐮¯|2dx+1r𝒬3r/4+|𝐮¯|2dxdt\displaystyle\sup_{\mathcal{I}_{3r/4}}\frac{1}{r}\int_{\mathcal{B}_{3r/4}^{+}}|\overline{\bf u}|^{2}\,\mathrm{d}x+\frac{1}{r}\int_{\mathcal{Q}_{3r/4}^{+}}|\nabla\overline{\bf u}|^{2}\,\mathrm{d}x\,\mathrm{d}t r3r(𝐮¯,π¯)+(r3r(𝐮¯,π¯))2/3.\displaystyle\lesssim r^{3}\mathscr{E}_{r}(\overline{\bf u},\overline{\pi})+\big{(}r^{3}\mathscr{E}_{r}(\overline{\bf u},\overline{\pi})\big{)}^{2/3}. (83)

Arguing as in (71) and (72) and appreciating the correct scaling (that is, we have |ζ|r1|\nabla\zeta|\lesssim r^{-1}, |tζ|r2|\partial_{t}\zeta|\lesssim r^{-2} and |2ζ|r2|\nabla^{2}\zeta|\lesssim r^{-2}), the additional terms

1r7/2rr+|𝐮¯|3/2dxdσand1r2rr+|ζ𝐮¯|3/2dxdσ\displaystyle\frac{1}{r^{7/2}}\int_{\mathcal{I}_{r}}\int_{\mathcal{B}^{+}_{r}}|\overline{\bf u}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma\quad\text{and}\quad\frac{1}{r^{2}}\int_{\mathcal{I}_{r}}\int_{\mathcal{B}^{+}_{r}}|\zeta\nabla\overline{\bf u}|^{3/2}\,\mathrm{d}x\,\mathrm{d}\sigma

appear in our situation. The first term can be estimated by

1r2rr+|𝐮¯|3dxdσ+1r5rr+dxdσr3r(𝐮¯,π¯)+1\displaystyle\frac{1}{r^{2}}\int_{\mathcal{I}_{r}}\int_{\mathcal{B}^{+}_{r}}|\overline{\bf u}|^{3}\,\mathrm{d}x\,\mathrm{d}\sigma+\frac{1}{r^{5}}\int_{\mathcal{I}_{r}}\int_{\mathcal{B}^{+}_{r}}\,\mathrm{d}x\,\mathrm{d}\sigma\lesssim r^{3}\mathscr{E}_{r}(\overline{\bf u},\overline{\pi})+1

and, similarly, we have the upper bound

κrrr+|ζ𝐮¯|2dxdσ+cκ\displaystyle\frac{\kappa}{r}\int_{\mathcal{I}_{r}}\int_{\mathcal{B}^{+}_{r}}|\zeta\nabla\overline{\bf u}|^{2}\,\mathrm{d}x\,\mathrm{d}\sigma+c_{\kappa}

with κ>0\kappa>0 arbitrary for the second one. Here the remaining integral can be absorbed in the left-hand side of (83) if κ\kappa is sufficiently small. In conclusion, we obtain an additional (additive) constant on the right-hand side of (83). We could allow a quantity which disappears in the limit r0r\rightarrow 0, but a nontrivial constant destroys the argument for the proof of (82).

6.2. Optimality of the assumptions on the boundary

We are able to significantly relax the assumptions regarding the regularity of the boundary made in [26] (C2C^{2} versus W21/p,pW^{2-1/p,p} for some p>154p>\frac{15}{4}). However, it is unclear if our assumptions in Theorems 2.6 and 2.7 are optimal or if they can still be weakened. The assumptions on the coefficients for the perturbed Stokes system in Lemma 4.2 are certainly optimal as the theory of Sobolev multipliers already yields optimal results for the Laplace equation, see [19, Chapter 14]. There are various stages in the proof of the blow up lemma (see Lemma 5.1), which require restrictions on the regularity of 𝚽{\boldsymbol{\Phi}} (and hence that of φ\varphi):

  • In order to control the local energy inequality in the flat geometry second derivatives of 𝚽m{\boldsymbol{\Phi}}_{m} appear (see (I)m2(I)_{m}^{2} in the proof of (76)). We need 𝚽m2,3/2(δ){\boldsymbol{\Phi}}_{m}\in\mathcal{M}^{2,3/2}(\delta) (and hence φm4/3,3/2(δ)\varphi_{m}\in\mathcal{M}^{4/3,3/2}(\delta)) to estimate the critical term. We do not expect that it is possible to relax this.

  • We need the limit function 𝐯¯\overline{\bf v} - solution to some perturbed Stokes system - to be Hölder-continuous. By parabolic embeddings this follows from an estimate in

    Lt5/3Wx2,pWt1,5/3Lxp,\displaystyle L^{5/3}_{t}W^{2,p}_{x}\cap W^{1,5/3}_{t}L^{p}_{x},

    which requires 𝚽m2,p(δ){\boldsymbol{\Phi}}_{m}\in\mathcal{M}^{2,p}(\delta) (and hence φm21/p,p(δ)\varphi_{m}\in\mathcal{M}^{2-1/p,p}(\delta)) provided p>154p>\frac{15}{4}. However, we believe that Hölder-continuity of 𝐯¯\overline{\bf v} can be proved under much weaker assumptions. Unfortunately, we were unable to trace a suitable reference. Of course, it will be absolutely necessary to have continuity of solutions of the perturbed Stokes system. The latter is used as a local comparison system and continuity of the solution to the perturbed Navier–Stokes system cannot be expected otherwise.

  • In order to control the decay of the pressure to arrive at the contradiction in the blow up lemma (see Lemma 5.1) we need an estimate for 𝔮~m\nabla\tilde{\mathfrak{q}}_{m} in Lt5/3LxpL^{5/3}_{t}L^{p}_{x} with some p>154p>\frac{15}{4}. For this, the same assumptions on 𝚽m{\boldsymbol{\Phi}}_{m} as in the previous bullet point are needed. Here, we expect that it will be very difficult to relax this. Note in particular, that the Stokes system does not allow for estimates on the pressure in spaces with differentiability less than 1 (as in this case the time derivative only exists as a distribution on the solenoidal test-functions). This estimate also motivated the choice of our excess functional for which the time-integrability of the pressure is chosen as large as possible, while it should still coincide with the space integrability (in fact, one could allow a smaller space integrability, but this does not seem to improve the estimate for the size of the singular set). The more customary choice

    rt0,x0(𝐮¯,π¯)\displaystyle{\mathscr{E}}_{r}^{t_{0},x_{0}}(\overline{\bf u},\overline{\pi}) :=𝒬r+(t0,x0)|𝐮¯|3dydσ+r3𝒬r+(t0,x0)|π¯(π¯)𝒬r+(x0)|32dydσ,\displaystyle:=\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{r}(t_{0},x_{0})}|\overline{\bf u}|^{3}\,\mathrm{d}y\,\mathrm{d}\sigma+r^{3}\mathop{\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int}_{\mathcal{Q}^{+}_{r}(t_{0},x_{0})}|\overline{\pi}-(\overline{\pi})_{\mathcal{Q}^{+}_{r}(x_{0})}|^{\frac{3}{2}}\,\mathrm{d}y\,\mathrm{d}\sigma,

    introduced initially in [18], requires the more restrictive assumption p>92p>\frac{9}{2}. The original excess functional from [6], which is based on the function space Lt5/4Lx5/3L_{t}^{5/4}L^{5/3}_{x} for the pressure, leads to the condition p>152p>\frac{15}{2} instead. In general, for πLtrLxs\pi\in L_{t}^{r_{\ast}}L^{s_{\ast}}_{x} one has the condition p>3r2r2p>\tfrac{3r_{\ast}}{2r_{\ast}-2}.

6.3. Fluid-structure interaction

In a typical problem from fluid-structure interaction an elastic structure is located at a non-trivial part ΓΩ\Gamma\subset\partial\Omega, where Ω\Omega plays the role of a reference geometry. The deformation of the structure is described by a function η:I×Γ\eta:I\times\Gamma\rightarrow\mathbb{R} and the domain Ω\Omega is deformed to Ωη(t)\Omega_{\eta(t)} defined through its boundary

Ωη(t)={y+η(t,y)ν(y):yΩ},\displaystyle\partial\Omega_{\eta(t)}=\{y+\eta(t,y)\nu(y):\,y\in\partial\Omega\},

where ν\nu is the outer unit normal at Ω\partial\Omega. The Navier–Stokes equations are posed in the moving space-time cylinder given by

I×Ωη:=tI{t}×Ωη(t)\displaystyle I\times\Omega_{\eta}:=\bigcup_{t\in I}\{t\}\times\Omega_{\eta(t)}

with some abuse of notation. The displacement η\eta is the solution to a hyperbolic equation, a proto-typical example is

ϱst2η+βΔ2η=𝐅inI×Γ\displaystyle\varrho_{s}\partial_{t}^{2}\eta+\beta\Delta^{2}\eta={\bf F}\quad\text{in}\quad I\times\Gamma (84)

with ϱs,β>0\varrho_{s},\beta>0 describing a linearised Koiter-type shell. The function 𝐅{\bf F} on the right-hand side describes the response of the structure to the surface forces of the fluid imposed by the Cauchy stress. The existence of a weak solution to the coupled system has been shown in [17]. Solutions belong to the energy space which means for the structure that

ηW1,(I;L2(Γ))L(I;W2,2(Γ)).\displaystyle\eta\in W^{1,\infty}(I;L^{2}(\Gamma))\cap L^{\infty}(I;W^{2,2}(\Gamma)).

The second function space (and Sobolev’s embedding in two dimensions) indicates that the boundary of Ωη(t)\Omega_{\eta(t)} is not even Lipschitz. With only this information at hand we do not expect that a theory similar to that in the present paper is reachable. It has, however, been shown very recently in [20] that solutions satisfy additionally

ηW1,2(I;Wθ,2(Γ))L2(I;W2+θ,2(Γ)).\displaystyle\eta\in W^{1,2}(I;W^{\theta,2}(\Gamma))\cap L^{2}(I;W^{2+\theta,2}(\Gamma)). (85)

for all θ<12\theta<\frac{1}{2}. Due to the compact embeddings

W2+θ,2(Γ)W1,(Γ),W2+θ,2(Γ)W21/p,p(Γ),\displaystyle W^{2+\theta,2}(\Gamma)\hookrightarrow\hookrightarrow W^{1,\infty}(\Gamma),\quad W^{2+\theta,2}(\Gamma)\hookrightarrow\hookrightarrow W^{2-1/p,p}(\Gamma),

from some p>154p>\tfrac{15}{4} we obtain Lip(Ωη(t))δ\mathrm{Lip}(\partial\Omega_{\eta(t)})\leq\delta and Ωη(t)21/p,p(2)(δ)\partial\Omega_{\eta(t)}\in\mathcal{M}^{2-1/p,p}(\mathbb{R}^{2})(\delta) as required in Theorems 2.6 and 2.7. Unfortunately, this smallness is not uniformly in time. While we expect that methods similarly to those developed here can also be applied to study Navier–Stokes equations in moving domains (this is already highly non-trivial and requires some additional research), it is not clear if the regularity from (85) will be sufficient. Some further analysis is required.

Compliance with Ethical Standards

Conflict of Interest. The author declares that he has no conflict of interest.

Data Availability. Data sharing is not applicable to this article as no datasets were generated or analysed during the current study.

References

  • [1] D. Albritton, E. Brué, M. Colombo: Non-uniqueness of Leray solutions of the forced Navier-Stokes equations. Ann. of Math. 196, 415–455. (2022)
  • [2] D. Breit: Existence theory for generalized Newtonian fluids. Mathematics in Science and Engineering. Elsevier/Academic Press, London. (2017)
  • [3] D. Breit: Regularity results in 2D fluid-structure interaction. Preprint at arXiv:2207.14159v2
  • [4] T. Buckmaster, V. Vicol: Nonuniqueness of weak solutions to the Navier-Stokes equation. Ann. of Math. 189, 10–144. (2019)
  • [5] T. Buckmaster, M. Colombo, V. Vicol: Wild solutions of the Navier-Stokes equations whose singular sets in time have Hausdorff dimension strictly less than 1. J. Eur. Math. Soc. 24, 3333–3378. (2021)
  • [6] L. Caffarelli, R. V. Kohn, L. Nirenberg: Partial regularity of suitable weak solutions of the Navier–Stokes equations. Comm. Pure Appl. Math. 35, 771–831. (1982)
  • [7] A. Cianchi, V. G. Maz’ya: Global boundedness of the gradient for a class of nonlinear elliptic systems, Arch. Rational Mech. Anal. 212, 129–177. (2014)
  • [8] A. Cianchi, V. G. Maz’ya: Second-Order Two-Sided Estimates in Nonlinear Elliptic Problems. Arch. Rational Mech. Anal. 229, 569–599. (2018)
  • [9] H. J. Choe, J. L. Lewis: On the singular set in the Navier–Stokes equations. J. Funct. Anal. 175, 348–369. (2000)
  • [10] W. Desch, M. Hieber, J. Prüss: LpL^{p}-theory of the Stokes equation in a half space. J. Evol. Equ. 1, 115–142. (2001)
  • [11] L. Escauriaza, G. Seregin, V. Šverak: L3,L_{3,\infty}-solutions to the Navier–Stokes equations and backward uniqueness. Usp. Mat. Nauk 58, 3-44. (2003)
  • [12] R. Farwig, H. Sohr: The stationary and nonstationary Stokes system in exterior domains with nonzero divergence and nonzero boundary data. Math. Meth. Appl. Sci., 17, 269–291. (1994)
  • [13] G. P. Galid: An Introduction to the Mathemaical Theory of the Navier–Stokes equations. Steady-Sate Problems. 2nd Edition. Springer Monographs in Mathematics. Springer, New York Dordrecht Heidelberg London. (2011)
  • [14] Hieber, Matthias; Saal, Jürgen: The Stokes equation in the LpL^{p}-setting: well-posedness and regularity properties. Handbook of mathematical analysis in mechanics of viscous fluids, 117–206, Springer, Cham. (2018)
  • [15] O. A. Ladyzhenskaya, G. A. Seregin: On Partial Regularity of Suitable Weak Solutions to the Three-Dimensional Navier–Stokes equations. J. Math. Fluid Mech. 1, 356–387. (1999)
  • [16] J. Leray: Sur le mouvement d’un liquide visqueux emplissant l’espace. Acta Math. 63, 193–248. (1934)
  • [17] D. Lengeler, M. Růžička; Weak Solutions for an Incompressible Newtonian Fluid Interacting with a Koiter Type Shell. Arch. Rational Mech. Anal. 211, 205–255. (2014)
  • [18] F.-H. Lin: A new proof of the Caffarelli–Kohn–Nirenberg theorem. Comm. Pure Appl. Math. 51, 241–257. (1998)
  • [19] V. G. Maz’ya, T. O. Shaposhnikova: Theory of Sobolev multipliers, volume 337 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin. With applications to differential and integral operators. (2009)
  • [20] B. Muha, S. Schwarzacher: Existence and regularity for weak solutions for a fluid interacting with a non-linear shell in 3D. Ann. I. H. Poincaré – AN, DOI: 10.4171/AIHPC/33
  • [21] T. Runst, W. Sickel: Sobolev Spaces of Fractional Order, Nemytskij Operators, and Nonlinear Partial Differential Equations. De Gruyter Series in Nonlinear Analysis and Applications, vol.3, Walter de Gruyter & Co., Berlin, New York. (1996)
  • [22] J. Saal: Stokes and Navier–Stokes equations with Robin boundary conditions in a halfspace. J. Math. Fluid Mech. 8, 211–241. (2006)
  • [23] G. A. Seregin: Local regularity of suitable weak solutions to the Navier–Stokes equations near the boundary, J. Math. Fluid Mech. 4, 1–29. (2002)
  • [24] G. A. Seregin: Some estimates near the boundary for solutions to the nonstationary linearized Navier–Stokes equations. Zap. Nauchn. Semin. POMI 271, 204–223. (2000)
  • [25] G. A. Seregin: Remarks on regularity of weak solutions to the Navier–Stokes system near the boundary. Zap. Nauchn. Semin. POMI, 295, 168–179. (2003)
  • [26] G. A. Seregin, T. N. Shilkin, V. A. Solonnikov: Boundary partial regularity for the Navier–Stokes equations. J. Math. Sci. (N.Y.) 132, 339–358. (2006)
  • [27] V. Scheffer: Hausdorff measure and the Navier–Stokes equations. Comm. Math. Phys. 55, 97–112. (1977)
  • [28] V. Scheffer: Partial regularity of solutions to the Navier–Stokes equations. Pacifc J. Math. 66, 535–552. (1976)
  • [29] V. Scheffer: The Navier–Stokes equations in a bounded domain. Comm. Math. Phys. 73, 1–42. (1980)
  • [30] V. Scheffer: Boundary Regularity for the Navier-Stokes Equations in a Half-Space. Comm. Math. Phys. 85, 275-299. (1982)
  • [31] V. A. Solonnikov: Estimates for Solutions of Nonstationary Navier-Stokes Equations. J. Soviet Math. 8, 467–528. (1977)
  • [32] V. A. Solonnikov: On estimates of solutions of the nonstationary Stokes problem in anisotropic S. L. Sobolev spaces and estimate of the resolvent of the Stokes problem. Usp. Mat. Nauk, 58, 123–156. (2003)
  • [33] H. Triebel, Theory of Function Spaces, Modern Birkhäuser Classics, Springer, Basel. (1983)
  • [34] H. Triebel, Theory of Function Spaces II, Modern Birkhäuser Classics, Springer, Basel. (1992)